You are on page 1of 196

Structure and Bonding  157

Series Editor: D.M.P. Mingos

Banglin Chen
Guodong Qian  Editors

Metal-Organic
Frameworks
for Photonics
Applications
157
Structure and Bonding

Series Editor:

D.M.P. Mingos, Oxford, United Kingdom

Editorial Board:

F.A. Armstrong, Oxford, United Kingdom


X. Duan, Beijing, China
L.H. Gade, Heidelberg, Germany
K.R. Poeppelmeier, Evanston, IL, USA
G. Parkin, NewYork, USA
M. Takano, Kyoto, Japan

For further volumes:


http://www.springer.com/series/430
Aims and Scope

The series Structure and Bonding publishes critical reviews on topics of research
concerned with chemical structure and bonding. The scope of the series spans the
entire Periodic Table and addresses structure and bonding issues associated with all of
the elements. It also focuses attention on new and developing areas of modern
structural and theoretical chemistry such as nanostructures, molecular electronics,
designed molecular solids, surfaces, metal clusters and supramolecular structures.
Physical and spectroscopic techniques used to determine, examine and model struc-
tures fall within the purview of Structure and Bonding to the extent that the focus is on
the scientific results obtained and not on specialist information concerning the
techniques themselves. Issues associated with the development of bonding models
and generalizations that illuminate the reactivity pathways and rates of chemical
processes are also relevant

The individual volumes in the series are thematic. The goal of each volume is to give
the reader, whether at a university or in industry, a comprehensive overview of an area
where new insights are emerging that are of interest to a larger scientific audience.
Thus each review within the volume critically surveys one aspect of that topic and
places it within the context of the volume as a whole. The most significant develop-
ments of the last 5 to 10 years should be presented using selected examples to illustrate
the principles discussed. A description of the physical basis of the experimental
techniques that have been used to provide the primary data may also be appropriate,
if it has not been covered in detail elsewhere. The coverage need not be exhaustive in
data, but should rather be conceptual, concentrating on the new principles being
developed that will allow the reader, who is not a specialist in the area covered, to
understand the data presented. Discussion of possible future research directions in the
area is welcomed.

Review articles for the individual volumes are invited by the volume editors.

In references Structure and Bonding is abbreviated Struct Bond and is cited as a


journal.
Banglin Chen Guodong Qian l

Editors

Metal-Organic Frameworks
for Photonics Applications

With contributions by
B. Chen X.-M. Chen Y. Cui S. Du P. Falcaro
l l l l l

S. Furukawa K. Hirai S. Kitagawa W. Lin


l l l l

R. Medishetty G. Qian J.J. Vittal H. Zhang


l l l l

J.-P. Zhang T. Zhang


l
Editors
Banglin Chen Guodong Qian
Department of Chemistry Dpt. of Materials Science & Engineering
University of Texas at San Antonio Zhejiang University
San Antonio Hangzhou
Texas People’s Republic of China
USA

ISSN 0081-5993 ISSN 1616-8550 (electronic)


ISBN 978-3-642-44966-6 ISBN 978-3-642-44967-3 (eBook)
DOI 10.1007/978-3-642-44967-3
Springer Heidelberg New York Dordrecht London
Library of Congress Control Number: 2014931514

# Springer-Verlag Berlin Heidelberg 2014


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of
the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or
information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed. Exempted from this legal reservation are brief excerpts
in connection with reviews or scholarly analysis or material supplied specifically for the purpose of being
entered and executed on a computer system, for exclusive use by the purchaser of the work. Duplication
of this publication or parts thereof is permitted only under the provisions of the Copyright Law of the
Publisher’s location, in its current version, and permission for use must always be obtained from
Springer. Permissions for use may be obtained through RightsLink at the Copyright Clearance Center.
Violations are liable to prosecution under the respective Copyright Law.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt
from the relevant protective laws and regulations and therefore free for general use.
While the advice and information in this book are believed to be true and accurate at the date of
publication, neither the authors nor the editors nor the publisher can accept any legal responsibility for
any errors or omissions that may be made. The publisher makes no warranty, express or implied, with
respect to the material contained herein.

Printed on acid-free paper

Springer is part of Springer Science+Business Media (www.springer.com)


Preface

Photonics, the science of light emission, transmission, modulation, manipulation,


and detection, are playing an essential role in many fields, including information
processing, communication technologies, military, and biomedicine. Compared to
electronics, photonics enable the transport and processing of information at the
speed of light. The present boom in optical fiber technology is an excellent example
of how much impact photonics has on the advancement of our society. Up to now,
various inorganic and organic materials have been widely developed for photonics
applications, including nonlinear optical materials, luminescent materials, laser
materials, photoactive materials, photoconductors, and so on. The research on the
photonics materials combines the applied research of chemistry, solid state physics,
materials science, and optical and computer engineering and will be the most
dominating forces in various fields of science and engineering.
Currently, metal-organic frameworks (MOFs) are being intensively studied as a
novel class of hybrid inorganic–organic material with ultrahigh porosity, enormous
internal surface areas, together with the extraordinary tailorability of structure,
dimension, size, and shape. Although the photonics application of MOF materials
is still at the early stage compared with the application for gas storage, separation,
and heterogeneous catalysis, the currently available results have unambiguously
demonstrated that the design and construction of MOFs for photonics functionality
is very active. Because of the inherent advantages of both organic links (easily
modify, flexibility, versatility, etc.) and inorganic metal ions (unique electronic and
optical nature), MOFs will open a land of promising applications in photonics fields
where conventional inorganic or organic materials might not be suitable. The MOF
approach can also offer a variety of other attractive characteristics such as the
straightforward syntheses, predictable structures and porosities, and collaborative
properties to develop new photonics materials and important applications.
Due to the importance as well as the rapid progress of the MOF materials in the
field of photonics, it is quite appropriate to publish this first book on MOFs for
photonics application. The book presents the most up-to-date source of information
on photonics applications of MOFs and not covered yet by other books. In this

v
vi Preface

book, the database of photonics MOFs, the design principles, the unique character-
istics of MOFs, and the potential photonics applications are discussed in detail. The
results obtained in the last years by the main leading MOF researchers all over the
world are presented in this book, thus providing a valid and precious overview on
the last developments and moreover on the future innovative applications of the
MOF materials in the field of photonics for scientists working in this area.
The book is organized in six chapters and opens with a chapter titled “Design
and Construction of Metal-Organic Frameworks” by Chen and Zhang, which
reviews the fundamental characteristics, molecular design, and synthetic strategies
of MOFs. The chapter titled “Luminescent Properties and Applications of Metal-
Organic Frameworks” by Qian et al. elaborates on the luminescent behavior of
MOFs and describes the applications of luminescent MOFs in the fields of lighting-
emitting, chemical sensors, and medicine. In chapter titled “Metal-Organic Frame-
works for Photocatalysis” by Lin discusses the photocatalysis application of MOF
materials. Vital et al. then describe the photochemical transformation within MOFs
in chapter titled “Photochemical Transformation within Metal-Organic Frame-
works”. The chapter titled “Metal-Organic Frameworks for Nonlinear Optics” by
Du et al. provides a comprehensive review of MOFs that display nonlinear optical
properties. Lastly, Furukawa and coworkers in chapter titled “Host-Guest Metal-
Organic Frameworks for Photonics” describe the luminescent properties of host-
guest MOFs and their potential applications as sensing materials.

San Antonio, TX, USA Banglin Chen


Hangzhou, People’s Republic of China Guodong Qian
Contents

Metal–Organic Frameworks: From Design to Materials . . . . . . . . . . . . . . . . . . . 1


Jie-Peng Zhang and Xiao-Ming Chen

Luminescent Properties and Applications of Metal-Organic


Frameworks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
Yuanjing Cui, Banglin Chen, and Guodong Qian

Metal-Organic Frameworks for Photocatalysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89


Teng Zhang and Wenbin Lin

Metal-Organic Frameworks for Photochemical Reactions . . . . . . . . . . . . . . . 105


Raghavender Medishetty and Jagadese J. Vittal

Metal–Organic Frameworks for Second-Order Nonlinear Optics . . . . . . . 145


Shaowu Du and Huabin Zhang

Host–Guest Metal–Organic Frameworks for Photonics . . . . . . . . . . . . . . . . . . 167


Kenji Hirai, Paolo Falcaro, Susumu Kitagawa, and Shuhei Furukawa

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

vii
Struct Bond (2014) 157: 1–26
DOI: 10.1007/430_2013_100
# Springer-Verlag Berlin Heidelberg 2013
Published online: 4 May 2013

Metal–Organic Frameworks: From Design


to Materials

Jie-Peng Zhang and Xiao-Ming Chen

Abstract Metal–organic frameworks are a new class of advanced porous materials,


which have several important characteristics, such as highly periodical, diverse and
designable structures, high porosity, unique and modifiable pore surface, as well as
framework flexibility. In this chapter, we give a brief account focusing on the
molecular design, synthetic strategies and interesting properties and applications
of metal–organic frameworks with a number of selected examples.

Keywords Application  Metal–organic framework  Molecular design  Porous


coordination polymer  Property  Synthesis

Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2 Design and Topology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1 Cluster-Based Frameworks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2 Framework Interpenetration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3 Synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
4 Properties and Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
4.1 Framework Flexibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4.2 Gas Storage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
4.3 Separation and Enrichment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.4 Molecular Sensing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

J.-P. Zhang (*) and X.-M. Chen


MOE Key Laboratory of Bioinorganic and Synthetic Chemistry, School of Chemistry and
Chemical Engineering, Sun Yat-Sen University, Guangzhou 510275, China
e-mail: zhangjp7@mail.sysu.edu.cn; cxm@mail.sysu.edu.cn
2 J.-P. Zhang and X.-M. Chen

1 Introduction

Although the term coordination polymer first appeared in 1960s, and can be traced up
to the first man-made coordination polymer or Prussian Blue by an accident in early
eighteenth century, the very extensive interest in coordination polymers was triggered
by several pioneering reports on discoveries of the porous structures and related
functionalities around 1990. In particular, Robson et al. reported the porous metal
cyanides with ion-exchange properties [1, 2], Kitagawa et al. demonstrated methane
adsorption in a microporous coordination polymer [3], and Yaghi et al. reported a rich
family of highly porous and thermally stable metal carboxylates [4] that were termed
metal–organic frameworks (MOFs) [5]. The research in coordination polymers has
grown substantially in the past decade, as being evidenced by the exponential increase
of publication number.
MOFs are constructed by metal ions/clusters and organic bridging molecules. They
can be regarded as a subclass of coordination polymers, since the metal ions/clusters
should be solely bridged by organic molecules (or ligands) via coordination into
two-dimensional (2D) or three-dimensional (3D) porous networks in MOFs [5],
whereas the bridging ligands may be either inorganic or organic in coordination
polymers [6]. Nevertheless, the term MOF has been very popularly used for different
types of coordination compounds with 3D to 1D, and even 0D, both porous and
nonporous structures. Anyway, we refer MOF to porous coordination polymer in
this chapter.
MOFs may have both advantages of inorganic and organic materials [7, 8], but their
physical and chemical stabilities are relatively lower than those of carbon materials
and inorganic zeolites solely composed of carbon–carbon and metal-oxide bonds,
respectively. As being well known that metal ions (or clusters) and organic bridging
ligands are diversified in coordination behavior and geometry, the combinations of
metal ions (or clusters) and bridging organic ligands through coordination interactions
are almost infinite. Consequently, a huge number of MOFs with different
functionalities, such as gas adsorption, separation, sensing, catalysis, and lumine-
scence, have been documented. MOFs have some important characteristics owing to
their unique compositions and structures, such as highly periodical structures, high
porosity, special and modifiable pore surface, as well as framework flexibility. The 3D
periodicity or crystallinity benefits the determination of their structures, in particular
by single-crystal X-ray diffraction, which is fundamental for a well understanding of
the structure–property relationship and better design of new materials. High porosity
can be rationally achieved by using suitable framework structure and/or long organic
ligands, which is important for adsorption and storage of molecules, catalysis, etc. The
pore surface of MOF is decorated by both inorganic moieties and various functional
groups of organic ligands, which can be readily tailored to meet a specific property or
function, such as selective adsorption, separation, and sensing. Compared to the
traditional zeolites and other inorganic porous materials, MOFs are very unique for
their highly flexible host framework structures, which are very useful for smart
materials with special sorption, storage, separation, and sensing properties [9].
Metal–Organic Frameworks: From Design to Materials 3

Therefore, as a new class of advanced porous materials, MOFs have become an


intense subject of scientific research. As many recent reviews on this field have been
published, we will give a brief account focusing on the molecular design, synthetic
strategies, and representative properties of MOFs by a series of selected examples.

2 Design and Topology

The bonding strength and directionality of coordination bonds are in between


covalent bonds and supramolecular interactions such as hydrogen bonds, which
imply that structural prediction of MOFs should be easier than that of a molecular
organic crystal. Although it is still difficult to emphasize the term design, there are
several strategies widely used to design the framework structures of MOFs.
In these strategies, the net-based approach or reticular chemistry is the most popular
and effective [10, 11]. As crystalline frameworks, the connectivities in inorganic
compounds, zeolites, and MOFs can be abstracted into periodical nets, or topologies,
constructed by nodes and linkers. Since the pioneering works of Wells, who introduced
the mathematical description for the structures of a variety of inorganic materials and
hypothetical networks [12], a large number of topologies have been known/predicted
to date, which can be achieved from open access databases such as the Reticular
Chemistry Structure Resource (RCSR) [13, 14]. The mathematical description of a
topology represents the inherent connectivity but is usually quite complicated and
difficult to index. For example, the net of carbon atoms, connected by carbon–carbon
covalent bonds in graphite and diamond have the point symbols 63 and 66, and vertex
symbols 6∙6∙6 and 62∙62∙62∙62∙62∙62, respectively. Therefore, topology symbols usually
borrow names from natural structures/compounds, such as honeycomb (for graphite)
and diamond. The RCSR symbols consisting of three bold letters, such as hcb and dia
(Fig. 1), were recommended as the uniform nomenclature [15], which cover the
IUPAC framework-type codes for all unique and confirmed zeolites (three capital
letters, such as SOD for sodalite) [16].
Network topologies have been used to not only simplify and describe but also
design the metal-ligand connectivities of MOFs. One can select metal ions/clusters
and/or organic ligands to mimic the node geometries of selected topologies (Fig. 2).
However, a given node geometry can be always suitable for construction of at least
several different topologies [17]. Therefore, simple and high-symmetry network
topologies are regarded as the most viable prototype of MOFs, because they reduce
the number of possible products of self-assembly. For example, combination of
two-coordinated coinage metal ions such as Cu(I) and Ag(I) with imidazolate
derivatives give 1D zigzag chains, helixes, or 0D rings. On the other hand, if
1,2,4-triazolates were used, coinage metal ions would tend to adopt trigonal-planar
coordination geometries, giving 3-connected topologies. The highly controllable
coordination mode and self-assembly behavior of azolate anions (such as imidazolate,
pyrazolate, and triazolate) make metal azolate frameworks (MAFs) a unique subset of
coordination polymers [18]. While more common ditopic neutral bridging ligands
such as dinitriles and diamines were used, coinage metal ions often showed tetrahedral
4 J.-P. Zhang and X.-M. Chen

a b

c d

Fig. 1 Selected simple topologies. (a) hcb, (b) dia, (c) sql, (d) pcu

N N N
NH NH NH
N
N N
O
O O O-
-
O
-
O
-
O O O-

N O- O O

N N N N

Fig. 2 Typical coordination geometries of metal ions and some simple bridging organic ligands

coordination, producing 4-connected dia networks. For divalent transition metal ions,
square-planar, tetrahedral, and octahedral coordination geometries are quite common,
which are suitable for the construction of sql, dia, and pcu topologies. The 2D sql
topology can be rationally constructed by rigid linear bipyridyl-type ligands and Ni(II)
or Cu(II) ions, in which the axial positions of the octahedral metal ions are occupied by
solvent molecules or small counter anions [19, 20]. On the other hand, pcu networks
based on single metal ions are relatively few because six pyridyl rings can induce
significant steric hindrance effect [21]. The dia topology may be the most frequently
encountered structures in MOFs because of the abundance of tetrahedral metal ions
Metal–Organic Frameworks: From Design to Materials 5

and linear ditopic bridging ligands [22]. Although some interesting porous compounds
are based on the common sql and dia topologies, low-dimensionality and/or inter-
penetration usually limit their porosities.
Based on the same square-planar node, the 3D nbo and lvt topologies are more
porous than the 2D sql one. However, mixing square-planar metal ions with linear
bridging ligands generally produce the 2D sql network rather than the two 3D
counterparts, because low-porosity structures are more energetically stable. Actually,
MOFs with the nbo and lvt topologies are very rare. Comparison of the three typical
topologies show that their difference relies on the dihedral angle between adjacent
square-planar nodes, in which they are coplanar, perpendicular, and with an acute
angle in sql, nbo, and lvt, respectively. To rationally construct the 3D 4-connected
networks based on square-planar nodes, control over the node-to-node orientation is
necessary. For example, Yaghi et al. rationally designed a highly porous MOF using
square-planar Cu2(RCOO)4 clusters and a bromine-substituted benzene-dicarboxylate
ligand, in which one of the carboxylate group is perpendicular to the benzene ring
because of the steric hindrance of the bromine atom [23]. With shorter bridging
ligands, one may gain more control over the network topologies. For example, we
have synthesized a series of Cu(I) 1,2,4-triazolates showing the sql, lvt, and nbo
topologies [24, 25]. In the first glance, these coordination networks can be simplified as
3-connected topologies because the metal ions and ligands are all 3-coordinated. But
as reflected by their RCSR symbols, sql-a, lvt-a, and nbo-a, they can be also described
as 4-connected topologies based on square-planar binuclear Cu2(Rtz)2 (tz denotes the
triazolate ring) nodes. Actually, sql-a, lvt-a, and nbo-a are obtained by 1,2,4-triazolate
derivatives with different substituent groups, in which the unsubstituted ligand give
the nonporous sql-a layer, the methyl substituted one give nonporous 2-fold inter-
penetrating lvt-a network, and larger substituents including ethyl and propyl groups
give porous non-interpenetrated nbo frameworks. In their crystal structures, steric
hindrance between pendant (non-coordinated) groups can be observed. It should be
noted that the relatively porous compound Cu(I) 3,5-diethyl-1,2,4-triazolate (MAF-2)
has demonstrated a number of interesting sorption properties by virtue of its unusual
flexibility on the coordination network and the side groups [26].

2.1 Cluster-Based Frameworks

Obviously, the easy rotation of molecular building blocks around individual coordi-
nation bonds is not beneficial for control over node-to-node orientation, which is very
important for construction of targeted topologies. Alternatively, rigid polynuclear
clusters such as Cu2(RCOO)4 and Cu2(Rtz)2 can provide more directionality; hence,
they can serve as secondary building units (SBUs) [5]. The most frequently used SBUs
are M2(RCOO)4L2, M4(μ4-O)(RCOO)6, and M3(μ3-O/OH)(RCOO)6L3. When L is a
monodentate terminal ligand and the carboxylate is a multitopic ligand, these SBUs
can serve as square-planar, octahedral, and trigonal-prismatic nodes, respectively
(Fig. 3). On the other hand, if L is a ditopic ligand, the dinuclear M2(RCOO)4 SBU
can serve as a linear (2-connected) linker [27], or a distorted octahedral (6-connected)
6 J.-P. Zhang and X.-M. Chen

O O O O
O O O O
L
O O O O
M M M M
O O O O O O O
M M M O M
L O O O O
L
O O O O
O O O O

O O
O OO O O OO O O OO O O
O M M L L
O M M M M M M
O O
O O O O O O O O O O O
M O O O O O O
M O O M O O M O O M O
O O O
O O O O O O
L

Fig. 3 The M2(RCOO)4, M4(μ4-O)(RCOO)6, and M3(μ3-O/OH)(RCOO)6 SBUs and their repre-
sentative coordination modes

a b c

Fig. 4 Structures of (a) HKUST-1, (b) MOF-5, and (c) MIL-101

node [28]. Similarly, the trinuclear M3(μ3-O/OH)(RCOO)6 SBU can serve as a


triangular (3-connected) node [29], or a tricapped trigonal-prismatic (9-connected)
node if L is ditopic [30, 31].
For example, Williams et al. reported that solvothermal reaction of Cu(NO3)2 and
benzene-1,3,5-tricarboxylic acid (H3btc) afforded [Cu3(btc)2(H2O)3] (HKUST-1)
with a highly symmetric porous network with a (3,4)-connected bor topology, in
which Cu2(RCOO)4(H2O)2 cluster and the btc3- ligand serve as the square-planar and
trigonal-planar nodes, respectively (Fig. 4a) [32]. Using expanded tricarboxylate
ligands, more porous analogs, sometimes interpenetrated, can be constructed [33, 34].
Yaghi et al. reported a series of pcu-type MOFs using the Zn4(μ4-O)(RCOO)6
cluster and linear dicarboxylate ligands with different lengths and side groups [35],
which are isoreticular with the prototypical structure [Zn4(μ4-O)(bdc)3] (MOF-5,
H2bdc ¼ benzene-1,4-dicarboxylic acid) (Fig. 4b). Several analogs of this
Metal–Organic Frameworks: From Design to Materials 7

oxo-centered tetranuclear cluster are also known. In a few examples, Zn(II) ions can
be fully or partially replaced by Co(II) ions, and the carboxylate groups can be also
replaced by pyrazolate groups [36]. Férey et al. developed various 6-connected
porous MOFs using the M3(μ3-O/OH)(RCOO)6L3 clusters. The trigonal-prismatic
geometry of M3(μ3-O/OH)(RCOO)6L3 SBU allows four such clusters to be arranged
in a supertetrahedron, which can further interconnect with each other by sharing their
vertexes to form zeolitic networks such as MIL-101 (Fig. 4c) [37]. For MOFs based
on the M2(RCOO)4L2 and M3(μ3-O/OH)(RCOO)6L3 SBUs, the monodentate termi-
nal ligand L can be removed to generate coordinatively unsaturated metal centers,
which can significantly enhance gas sorption affinities. It should be noted that the
bonding direction of the carboxylate groups in the M3(μ3-O/OH)(RCOO)6L3 can
change significantly, giving highly flexible frameworks structures when the network
topology is suitable [38].
Besides the carboxylate sites, the monodentate coordination sites (or the equatorial
sites) can coordinate not only to terminal ligands but also to bipyridyl-type bridging
ligands. Generally, the M2(RCOO)4L2 clusters form 2D sql layers with linear
dicarboxylates and different 0D polyhedra with bent dicarboxylates possessing differ-
ent bridging angles. When linear bipyridyl-type ligands are added, the 2D sql
M2(bdc)2L2 layers can be connected in both sides to form 3D unsymmetrical pcu
networks with a general composition of M2(bdc)2(bpy)2 (bpy ¼ 4,40 -bipyridyl-type
ligands). On the other hand, only the terminal ligands on the outer side of the polygons
can be substituted by bridging ligands. In this context, the dinuclear SBUs can be
regarded as 5-connected nodes (Fig. 3). On the other hand, the polyhedron can be
viewed as a super SBU [39]. For example, the M2(RCOO)4L2 SBUs can be linked as
cubooctahedra and octahedra using bent dicarboxylates with 120 and 90 linking
angles, respectively [40, 41], which can be further connected by bipyridyl-type ligands
to form fcu and pcu-type 3D frameworks considering the polyhedra as 12- and
6-connected nodes [39, 42–44]. In this type of porous networks, small cavities are
connected to the large channels via the small apertures on the polyhedra.
With three additional connections from the equatorial plane, the M3(μ3-O/OH)
(RCOO)6L3 SBUs can serve as tricapped trigonal-prismatic 9-connected nodes. Using
tritopic pyridyldicarboxylate ligands can fulfill the 2:1 molar ratio requirement of
carboxylate and pyridyl groups of the trinuclear clusters, which produce a new type of
(3,9)-connected topology namely xmz (Fig. 5a) [45, 46]. Interestingly, changing the
ratio of bridging lengths of the carboxylate and pyridyl ends of the tripodal ligands can
drastically alter the bonding direction of the carboxylate groups of the clusters and
even generate drastic framework breathing behaviors [47]. Combination of
dicarboxylate and pyridylcarboxylate ligands in a 1:2 ratio can also satisfy the
coordination requirement of the trinuclear clusters, which give rise to a series of
uninodal 9-connected networks with the ncb topology (Fig. 5b). The coordination
geometries of the clusters in the ncb networks are also influenced by the bridging
lengths of the ligands. Therefore, the ncb-type MOFs can be only synthesized by
dicarboxylate-pyridylcarboxylate ligand combinations with suitable bridging length
ratios, which was confirmed by geometry calculations and a systematic synthetic trial
using a large number of ligand combinations with different bridging lengths [30, 31].
8 J.-P. Zhang and X.-M. Chen

a b

Fig. 5 The (a) xmz and (b) ncb topologies

It should be noted that, even when the bridging ligands are very long, no inter-
penetration was observed for the ncb-type MOFs.

2.2 Framework Interpenetration

Interpenetration is a common phenomenon for coordination polymers, which is


usually regarded as a disadvantage for construction of highly porous frameworks.
But in some cases, it can be used to enhance framework stability and increase gas
sorption ability (at low pressures) [48–50], or generate interesting properties such as
stimuli responsive behaviors [51]. Nevertheless, there are several strategies to avoid
framework interpenetration. It is interesting that the nbo network seems to prevent
network interpenetration because its dual net (i.e., its channel system) is 8-connected
bcu rather than itself [23]. In other words, interpenetration often occurs for self-dual
topologies such as dia because crystals tend to contain identical, repeating units.
Actually, the non-interpenetrating ncb frameworks can be also simplified as
bcu networks with nbo channels regarding the triakistetrahedral supercages as
8-conncted nodes [30, 31]. However, some interpenetrating nbo networks have been
observed since 2003 [52, 53]. Due to imperfect interpenetration, interpenetrating
nbo networks are generally porous [54]. Although the self-dual principle refers to
perfect interpenetration only, it is still very useful for the selection of topology
targets. Compared with the dia topology, the more complicated zeolitic topologies
are also based on tetrahedral nodes but can hardly show interpenetration. Consider-
ing the paramount importance of natural and synthetic zeolites, zeolitic and zeolite-
like MOFs have received intensive attention. There are hundreds of known topology
types for zeolites [16].
Imidazolate is an exobidentate ligand with bridging angle close to those observed in
zeolites. Therefore, it is ideal to combine imidazolate and tetrahedral transition metal
ions for zeolitic networks. Such 4-connected coordination networks, such as Zn(II) and
Metal–Organic Frameworks: From Design to Materials 9

Co(II) imidazolate, have been known for more than 30 years, but they possess neither
the zeolitic topology nor porous structures [55, 56]. This situation is just similar the
examples mentioned above, i.e., using molecular building blocks mimicking the
topology geometry is not enough to construct the desired topology. More in-depth
understanding of the topology features should be necessary. By introducing template
molecules and/or structure directing agents into the solvothermal reaction system, You
and coworkers first synthesized a series of Co(II) imidazolate frameworks with zeolite-
like topology [57, 58]. After removal of the guest molecules, most of these structures
collapse, this means that these host frameworks with unusual topologies/structures
must be stabilized by guest inclusion.
Again, the key for producing subtle structure difference is to use non-coordinated
side groups to control the orientation of the imidazolate linkers, which has been
comprehensively demonstrated by a large group of low-dimensional coinage metal
imidazolates [59–62]. Although a substituent group is generally considered as dis-
advantageous for generation of porosity, the first porous zeolitic MOF, namely
MAF-3, was constructed by Zn(II) and benzimidazolate, which possesses a distorted
SOD topology and a porosity of 18% [63]. While the very bulky phenyl ring
significantly reduces the porosity and aperture size, it is useful for molecular sieving
applications, especially when fabricated as membranes [64]. More porous and regular
zeolitic metal imidazolate frameworks, i.e., SOD-[Zn(mim)2] (MAF-4, Hmim ¼
2-methylimidazole), ANA-[Zn(eim)2] (MAF-5, Heim ¼ 2-ethylimidazole) and
RHO-[Zn(eim)2] (MAF-6) were subsequently obtained by using simple alkyl-
imidazolates (Fig. 6) [65, 66]. Due to its high porosity, high thermal and chemical
stability, and unique pore structure, MAF-4 has served as a prototypical MOF in many
studies [18]. Yaghi and coworkers then introduced the high-throughput solvothermal
method [67], which enabled the discovery of a variety of new zeolitic and zeolite-like
metal imidazolate frameworks (popularly known as zeolitic imidazolate frameworks,
ZIFs) with interesting structures and properties [68]. The substituent groups play a
very important role in the formation and stability of these zeolitic frameworks,
which could be further demonstrated by the successful synthesis of an analogous
structure of MAF-4, i.e., SOD-[Zn(mtz)2] (MAF-7, Hmtz ¼ 3-methyl-1,2,4-
triazolate) [69]. Generally, 1,2,4-triazolates tend to use all of the three nitrogen
atoms for coordination [70]. But in the case of MAF-7, the methyl group reinforces
the formation of an SOD network, leaving an uncoordinated nitrogen atom for
guest binding.
There are other strategies for rational construction of non-interpenetrated porous
frameworks. For 3D networks, it is obvious that interpenetration can only occur
when the porosity of a single network is larger than 50%. Therefore, one may
introduce small side groups to reduce the porosity of a single network and yet
increase the porosity of the final crystal. Of course, the crystal porosity obtained by
this method has an obvious upper limit. Considering that interpenetration of 3D
networks must occur in all three dimensions, partial blocking of the coordination
network can prevent interpenetration and retain high porosity (Fig. 7).
The pillared-layer structures have been rather widely reported for MOFs, which are
not interpenetrated because their relatively dense layers block one of the three
dimensions. Based on a suitable layer structure containing coordinatively unsaturated
10 J.-P. Zhang and X.-M. Chen

Fig. 6 Framework structures of (a) MAF-4, (b) MAF-5, and (c) MAF-6
Metal–Organic Frameworks: From Design to Materials 11

Fig. 7 Control of interpenetration of pcu-type structures. (a) Normal interpenetration, (b) a


formal pillared-layer structure consisting of dense layers and long pillars, and (c) a
non-interpenetrating pcu net consisting of porous layers and short pillars

sites, one can change the pillar ligand to obtain MOFs with different porosity and
different pore size and surface. The arrangement of pillar ligands is determined by
distribution manner of coordinatively unsaturated metal ions on the layer. For exam-
ple, Kitagawa et al. reported a series of pillared-layer MOFs with a general formula
[Cu2(pzdc)2L] (CPL-n, H2pzdc ¼ pyrizine-2,3-dicarboxylic acid), in which Cu(II)
ions are four-coordinated and bridged by pyrazinedicarboxylate ligands to form dense
layers, and these linear bipyridyl-type pillar ligands coordinate with the remained
coordination site of Cu(II) and connect the layers into a 3D porous framework. Since
the Cu(II) ions on the layer form relatively dense 1D arrays, CPL-n possess 1D
channels [3].
There are some porous MOFs constructed by porous layers and short pillars,
which can be categorized as informal pillared-layer structures which use the short
pillars to block two of the three dimensions. For example, when the bipyridyl ligands
are relatively long in the [M2(bdc)2(bpy)] frameworks with unsymmetrical pcu
topologies, interpenetration usually occurs [71]. On the other hand, if the pillar
ligand is short enough, such as for 1,4-diazabicyclo[2.2.2]octane, no interpenetration
can be observed even if the dicarboxylate ligand is very long [72].
The term pillared-layer implies that the 3D structure is hierarchical with the
layer formed first and the pillar connects later, but such phenomena can be hardly
observed directly since the self-assembly and crystallization of MOFs are generally
carried out in a one-pot reaction. Nevertheless, using suitable non-bridging ligands
instead of the pillar ligands, the individual single layer structure can be sometimes
synthesized. In an occasional case, the hypothetic monolayer and pillared double
layer have been isolated during the synthesis of a pillared-layer MOF [73].
In the context of synthesis, the reality is always more complicated than a simple
theoretical expectation. The molecular assembly processes are usually influenced by
a variety of factors, including the reaction parameters and minor differences in the
structures and properties of the metal ion/clusters and the ligands [74]. Therefore,
there are still a lot of compounds were unintentionally generated rather than
predesigned. However, many compounds have been successfully prepared by the
net-based approach or reticular chemistry strategy with either single metal ions or
metal clusters as nodes, as well as the pillar-layered strategy. With these strategies,
one can construct MOFs with different pore sizes, shapes and surface properties by
12 J.-P. Zhang and X.-M. Chen

using similar bridging ligands with different lengths and side groups. Moreover,
mixed ligands with the same coordination mode and length but different side groups
or cores can be used to design solid-solution-type frameworks (periodic lattice with
non-periodic metal/ligand distribution), which are emerging as an effective strategy
for pore surface tailoring [69, 75, 76].

3 Synthesis

To allow the self-assembly of metal ions and organic ligands, as well as crystallization,
MOFs are mostly synthesized by solution-based methods at different conditions. In the
simplest approach, MOF crystals automatically grow in a solution containing the metal
salt and organic ligand. When there are solubility problems for the reactants and/or
products, other approaches such as diffusion and hydrothermal/solvothermal reactions
may be useful, in particular, for growth of large and high-quality crystals for single-
crystal diffraction. The reaction method and reaction condition are crucial for the
construction of some types of MOFs, especially those containing polynuclear SBUs.
For example, solvothermal reactions with amide-based solvents have been widely
demonstrated to be very useful for the synthesis of MOFs based on M4(μ4-O)(RCOO)6
and other types of clusters.
Although some MOF structures were determined by powder diffraction techniques,
single crystal specimens with diameter ca. 100 μm and above are usually necessary for
structure elucidation of new MOFs. Therefore, the crystallization should be slowed
down and undisturbed. On the other hand, to synthesized bulk sample of known
structures, the reaction time and scale rather than crystal size are more important.
The reaction time of MOF synthesis can be greatly accelerated by increasing reactant
concentration, rapid mixing of reactants, heating, microwave irradiation, etc.
[77]. Some MOFs can be also rapidly synthesized by mechanochemical reaction of
the corresponding metal salt and organic ligand in the presence of very small amount
or even without solvent. Nevertheless, soluble metal salts are generally necessary for
mechanochemical reactions [78], which produce acid as the by-product and the
obtained MOF samples still require solvent washing. In exceptional cases, metal
oxide or hydroxide can be used as the reactants, which only generate water as a
by-product. For example, some metal imidazolate frameworks have been obtained
by solvent/additive-assisted mechanochemical reaction between stoichiometric ZnO
and imidazoles at room temperature. A few porous/zeolitic MAFs can be even
synthesized without addition of any solvent or additive by simply heating the mixture
of ZnO/Zn(OH)2 and the azole ligand [79]. In principle, these mechanochemical and
solvent-free reaction methods can greatly increase the space-time-yield and beneficial
for large-scale production, but difficult to ensure the product purity because impurity is
always contained by the reactants. Alternatively, higher-purity MAFs can be obtained
by aqueous ammonia-assisted reactions, in which ammonia increases the solubility of
metal oxides and helps the deprotonation of weakly acidic azoles [80].
While many structure design strategies have been developed, none of which can
guarantee that there is only one possible self-assembly result with a given set of
reactants. Note that solvent and counter ions always participate in the reaction
Metal–Organic Frameworks: From Design to Materials 13

system and often involve in the final products. When the chemical composition of
the coordination network is fixed, the obtained product would be limited to a few
supramolecular isomers. Energetically, porous networks are generally less stable
than nonporous ones. Therefore, as-synthesized MOF crystals always contain
solvent and/or template molecules, which help stabilizing the porous structures.
For MOFs containing large channels and small guest molecules, it is not easy to
judge if the guest molecules serve as templates.
In some cases, the template effect is crucial and can be clearly demonstrated
by the crystal structures. For example, Kitagawa et al. showed that reaction of
Zn(NO3)2, Li(TCNQ) (TCNQ ¼ 7,7,8,8-tetracyano-p-quinodimethane) and bpy in
pure methanol and mixed methanol/benzene gave two MOFs with very different
structures, namely [Zn(TCNQ)(bpy)]·6MeOH and [Zn(dTCNQ)(bpy)]·1.5benzene
[81, 82]. The TCNQ– anion disproportionates and dimerizes to form the TCNQ2
and dTCNQ2 dianions without and with benzene addition, respectively (Fig. 8).
These dianion ligands contain excess electrons which can form strong charge
transfer-type interactions with aromatic molecules. The benzene molecules are
embedded in the small cavities composed of the dTCNQ2 ligands. In this context,
the template molecule can even influence the redox equilibrium of the reaction
system, resulting in unusual in situ ligand transformation during MOF synthesis.
The in situ ligand reactions have been observed more frequently during
solvothermal reactions, in which the high reaction temperature and catalytic action
of metal ions may promote organic reactions that are infeasible at common conditions.
Actually, many in-situ metal/ligand reactions have been reported to be useful for
synthesis of coordination polymers, such as hydrolysis of carboxylate esters, organic
nitriles, and aldehydes into the corresponding carboxylates, cleavage of acetonitrile/
ethylene carbon–carbon bonds and 1,3,4-oxadiazole carbon–nitrogen/carbon–oxygen
bonds, cleavage and formation of disulfide bonds, substitution of aromatic groups,
decarboxylation of aromatic carboxylates, cycloaddition of organic nitriles with azide
and ammonia, and so on (Fig. 8) [83, 84]. For example, in situ solvothermal generation
of carboxylate groups by hydrolysis of nitrile or ester groups is an effective synthetic
method to obtain the non-centrosymmetrical (or nonlinear optical active) coordination
polymers, which are not easily formed by conventional reactions of carboxylic acids
and metal ions [22]. In situ solvothermal reactions of iron powder with organic
carboxylic acid are also critical for the syntheses of magnetically interesting Fe(II)
and mixed-valence Fe(II,III) coordination polymers [85, 86], and some mixed-ligand
coordination polymers, which are hardly to be prepared by direct mixing of the related
ligand and metal ions, could possibly be achieved by decomposition of an organic
molecule into two different bridging ligands [87].
It is also noteworthy that MOFs, although generally regarded to be much instable
than conventional adsorbents, can undergo covalent post-synthetic modification or
metal/ligand exchange [88, 89]. Nevertheless, hydrothermal/solvothermal reactions
are important for searching new, interesting MOFs and growing large crystals for
structural characterization by precise single-crystal X-ray diffraction. However,
for practical applications, large-scale and rapid synthesis is also a very important
objective. Therefore, conventional reactions by direct mixing of the reactants, as well
as microwave-assistant and mechanochemical reactions are very attractive.
14 J.-P. Zhang and X.-M. Chen

NC CN
2- - -
2
NC CN NC CN NC CN

NC CN
+
NC CN

NC CN NC CN NC CN

NC CN

R N
N
N3- N NH
H
R N R
R N+ NH3
N N
NH2
N2H4 R N R
N N

N N

N N
N N
HO H
N N
+

N
N
H OH

N N H
H
N N

N N
H OH

Fig. 8 Representative in situ organic reactions occurred during MOF synthesis

4 Properties and Applications

By virtue of the 3D periodic structures, the porous structures of MOFs can be straight-
forwardly visualized from the crystal structures, which is very important for under-
standing the structure–property relationship. However, the host framework may not
Metal–Organic Frameworks: From Design to Materials 15

retain after guest removal, because coordination bonds are relatively weak compared
with covalent bonds in traditional adsorbents. Many MOFs collapse after guest
removal and heat treatment, meaning that there are only pseudo or virtual porosity in
the as-synthesized crystals. Therefore, rigid frameworks with high thermal stability
have been the main research target [90]. So far, many MOFs can be stable up to
300–400 C and a few to 500 C, which are high enough for many applications.
Actually, the chemical stability of MOFs is a more serious problem. Most
carboxylate-based MOFs are unstable in water or moist air. In contrast, MAFs
especially those based on diazolates (imidazolates and pyrazolates) have demonstrated
high thermal and chemical stabilities, which can be attributed to their relatively strong
metal-azolate coordination bonds and chemically inert pore surfaces.
To confirm the permanent porosity of the sample, a gas sorption isotherm is
necessary, which is commonly carried out by nitrogen at 77 K. Argon with smaller
size and spherical molecular shape can measure pore surface area and pore volume
more precisely. When the pore size is too small to be entered by nitrogen, carbon
dioxide and hydrogen with smaller kinetic diameters can be used instead. The gas
sorption behavior is determined by the adsorbate, the temperature and pressure, and the
structural characteristics of adsorbent such as pore size and shape, surface area, and
pore volume. By virtue of the diversified and unique structures, MOFs have
demonstrated many interesting properties potentially useful for practical applications.

4.1 Framework Flexibility

While rigid frameworks are associated with high stability and usefulness for practical
applications, many MOFs can change their structures under external stimuli, because
their host frameworks consist of relatively weak bonds, rotatable single bonds, weak
supramolecular interaction between multiple networks, etc. [9, 91, 92]. For example,
even for the seemingly very rigid structure MOF-5, the phenyl ring of the bdc2 ligand
is rotatable [93], although the overall crystal lattice is not distorted. More commonly,
rotation of molecular building blocks around the metal-ligand bonds or single bonds
in the organic ligands can give rise to framework distortion and crystal symmetry
change. Flexible coordination networks can also lose the long-range order to form
amorphous or quasi-amorphous phases. While MOFs with the pseudo-porosity which
irreversibly turn amorphous after guest removal, flexible MOFs show reversible
structural transformations either in a crystal to crystal or in a crystal to amorphous
manner. For example, 3,30 ,5,50 -tetramethyl-4,40 -bipyrazolate (bpz2) has been used
as flexible ditopic ligand to link trinuclear Ag3 and Cu3 clusters as 3-connected porous
frameworks [94–96]. The two pyrazolate rings are rotatable around their connection
(a C–C single bond), which possess a dihedral angle freely changeable between 50
and 90 . Therefore, the porous MOF [Ag2(bpz)] demonstrated a variety of crystal
symmetry after inclusion of different guests. In specific guest loading, a single-crystal
specimen may lose its diffraction ability although its shape and transparency are not
16 J.-P. Zhang and X.-M. Chen

significantly changed, indicating the formation of a highly distorted, non-periodic


network [95, 96].
To monitor the structure and crystallinity change of MOFs, powder X-ray
diffraction is a convenient technique. More precise host–guest structural information
could be obtained by synchrotron diffraction and single-crystal X-ray diffraction [97,
98]. Combination of other in-situ spectroscopic techniques can give more insightful
structural information [99], especially for non-periodic structural alterations.
Flexible MOFs can show very different types of framework flexibilities such as
gliding of 2D layers [19], shifting of interpenetrating 3D networks [100], huge
breathing of a single 3D network due to changeable bonding direction of organic
ligands [38]. The structural transformation of host framework is always associated
with energy change, which is generally induced by chemical stimuli, i.e., the
sorption, desorption, and exchange of guest molecules. This common kind of
framework flexibility usually shows abnormal guest sorption isotherms. Framework
flexibility induced by physical stimuli such as light and heat has been rarely reported.
For example, MAF-2 processes an nbo-a-type host framework, which defines a bcu-
type channel system with large cavities interconnected by small apertures. Interest-
ingly, the apertures are blocked by the obviously flexible ethyl groups, and N2
adsorption was observed at 195 K rather than at 77 K. Further, single-crystal
structures of the guest-free and N2 loaded phases showed that there is only one
thermodynamically stable state for the host framework, indicating that the ethyl
group is flexible with very low energy change and energy barrier, so that its motion is
determined mainly by heat instead of guest. After adsorption of different solvent
molecules, MAF-2 can also distort its metal triazolate framework to form cubic,
trigonal, and triclinic crystal symmetry (Fig. 9) [26]. Obviously, the ability of
flexible MOFs to change their host structures by chemical and physical stimuli is a
unique characteristic compared with conventional adsorbents, which may be used to
produce novel and useful properties.

4.2 Gas Storage

With appropriate framework design, many MOFs possess surface area and pore
volume much larger than conventional adsorbents such as activated carbons and
zeolites, which imply that MOFs are promising materials for gas storage. Never-
theless, the adsorbent is potentially useful only when its adsorption amount is larger
than that of an empty container. For example, MOFs with large pore volumes can
always adsorb very large amounts of nitrogen gas at 77 K, but it is not useful
because an empty container can store more liquid nitrogen at 77 K, or a gas cylinder
working at above 100 bar can easily store much gaseous nitrogen at room tempera-
ture. Similarly, MOFs may adsorb more CO2 than an empty cylinder at room
temperature at the low pressures, but they are not considered as good CO2 storage
medium because CO2 can be liquefied easily without excess pressurization.
Metal–Organic Frameworks: From Design to Materials 17

a b

c d

Fig. 9 Two kinds of framework flexibility observed in MAF-2. (a, b) Temperature-induced


closed and open states of the ethyl blocked aperture. (c, d) Guest-induced distortion of the
Cu-triazolate framework

Therefore, MOFs are generally considered as the storage medium for gases such as
hydrogen and methane that are difficult to compress or liquefy.
Hydrogen is a clean and high-energy fuel, but it is very difficult to store or
deliver because of its extremely low boiling point. Various strategies have been
studied for hydrogen storage, but none has met the practical requirements yet. For
example, chemical absorption can provide very large storage capacity but hydrogen
can be hardly released. Alternatively, physical adsorption of hydrogen in porous
materials such as MOFs is very fast [101, 102]. By virtue of the ultrahigh surface
areas and pore volumes, some MOFs have achieved the DOE (Department of
Energy, U.S.A.) targets for hydrogen storage, i.e., 6.5 wt% and 45 g L1. Note
that the density of liquid hydrogen is only ca. 70 g L1. However, these data can be
only obtained at very low temperatures such as 77 K, because the adsorption
enthalpy of physical adsorption is too low. There are different strategies to enhance
the adsorption affinity of MOFs. Pore size is a critical parameter that determines the
18 J.-P. Zhang and X.-M. Chen

adsorption affinity. However, very small pore size is not likely compatible with the
very large surface area and pore volume required for practical applications.
Incorporation of coordinatively unsaturated metal centers on the pore surface can
also effectively enhance adsorption affinity, but high-concentration metal ions
would generally increase the framework density.
As a main component in natural gases, methane is also an important fuel. The
boiling point of methane is very low but much higher than that of hydrogen.
Therefore, methane can be easier adsorbed by porous materials. A few MOFs
have demonstrated large methane storage capacity exceeded the DOE target, i.e.,
180 v/v at room temperature and 35 bar. For example, Zhou and coworkers
designed and synthesized a highly porous coordination network PCN-14, in
which the Cu2(RCOO)4 clusters are connected by anthracene tetracarboxylate
ligands into an nbo topology [103]. PCN-14 possesses a large Langmuir surface
area over 2,100 m2 g1 and a large pore volume over 0.8 cm3 g1, as well as
nanocage-type pore and large aromatic rings on the pore surface, which give rise to
methane uptake of 230 v/v and enthalpy of 30 kJ mol1 at 290 K and 35 bar.
While increasing the adsorption enthalpy or affinity has been the main strategy for
improving the gas storage performance, storage refers to not only deposit but also
withdraw, i.e., adsorption and desorption of an adsorbate. Actually, the effective
storage capacity equals to the difference of adsorption amounts at two temperature/
pressure points of interest rather than the adsorption amount at a single temperature/
pressure point. When the adsorption enthalpy is too high, desorption becomes
difficult although adsorption amount is high. It has been calculated that, assuming
a type-I hydrogen adsorption isotherm working between 1.5 and 35 bar at 298 K for a
typical storage system, the maximum storage capacity can be obtained by an
adsorption enthalpy of 15.1 kJ mol-1[104]. Actually, the type-I isotherm is very
steep at the lowest pressure ranges, which means very large amounts of adsorbates
cannot be desorbed. For example, the highly porous MOF-177 has not only a very
high hydrogen uptake of 48.3 g L1 at 77 K and 70 atm, but also a relatively high
uptake of 8 g L1 at 1.5 bar, meaning that its useful storage capacity is about 40 g L1
[105].
In special cases, the concept of useful storage capacity is more crucial. For
example, acetylene is highly unstable, and pure acetylene cannot be compressed
above 2 bar. Therefore, gas cylinder can only provide very limited storage capacity
because the working pressure difference is small. Although some microporous
adsorbents such as zeolites and MOFs can adsorb much more acetylene than for
an empty gas cylinder [106], their adsorption affinity is too strong, which prevent
acetylene desorption above 1 bar. While a normal type-I sorption isotherm tends to
retain much guest at low pressure, a type III or V isotherm can provide more useful
storage capacity, but these isotherms are generally associated with low saturated
adsorption amounts and weak adsorption affinity (Fig. 10). Therefore, the adsorp-
tion affinity should be lowered or the isotherm shape should be modified to
maximize the uptake difference between the lower and upper working temperature
for pure acetylene, which can be achieved by an MOF with suitable pore surface
structure and framework flexibility [107].
Metal–Organic Frameworks: From Design to Materials 19

Fig. 10 Comparison of the usable storage capacity of Type I, III, and V isotherms

4.3 Separation and Enrichment

The sorption behaviors of different adsorbates are always different, which can be used
to separate or enrich one of the adsorbates from the mixtures [108, 109]. Unlike
storage, separation does not require large surface area and pore volume. Alternatively,
small pore size is commonly beneficial for high separation performance. Adsorbents
can separate different adsorbates in several mechanisms. The most effective one is the
molecular sieving effect, which only adsorbs the smaller molecules and excludes the
larger molecules completely, giving very high separation ratio. Crystalline MOFs are
highly ordered frameworks as for zeolites, which can be also engineered to have
suitable pore sizes and shapes [110]. Further, flexible MOFs can selectively change
their pore sizes. To utilize the molecular sieving properties, adsorbents may be used in
different manners. The most straightforward and economic method is to fabricate an
adsorbent-based membrane. Gas sieving membranes have been successfully
fabricated with several types of MOFs such as MAF-3, MAF-4, and HKUST-1
[64, 111, 112].
More commonly, two kinds of molecules are separated by the different adsorption
affinity or adsorption speed. For example, carbon dioxide is regarded as a main
source of green house effect, and separation of carbon dioxide from fuel gases has
attracted intensive attention recently [113]. Although chemical absorption of carbon
dioxide by amine-based solvents is very effective, the regeneration process requires
much energy, and the volatile, corrosive amine vapor is also a serious problem.
Physical adsorption is a viable approach to capture carbon dioxide. The boiling point
and quadruple moment of carbon dioxide is much higher than those of nitrogen and
oxygen in fuel gasses, which means carbon dioxide can be selectively adsorbed.
20 J.-P. Zhang and X.-M. Chen

However, the host–guest interaction between carbon dioxide and MOFs are gener-
ally too weak to effectively capture the low-concentration carbon dioxide. Never-
theless, this problem can be very likely solved by introducing coordinatively
unsaturated metal centers, amine groups, etc. on the pore surfaces. Of course, high
adsorption enthalpy will always produce desorption problems, which may be solved
by using multiple weak interactions [114, 115].
Generally, single-component sorption isotherms can be used to judge the potential
separation performance. The mixed-component sorption behavior is not a simple
mathematical addition of several single-component sorption isotherms. With only
single-component isotherms, the ideal adsorption solution theory (IAST) can be
applied to predict the mixture separation performance. The sorption kinetics is also
very crucial for separation. For example, the sorption isotherms of propane and
propene are very similar in MAF-4, but they can be effectively separated according
to their different diffusion speed [116]. Therefore, testing the separation performance
in more practical conditions is always desirable. Generally, the practical separation
performance of an MOF can be tested by the breakthrough experiment, in which the
mixture passes through the MOF filled column, and the time-dependent concentrations
of adsorbates are recorded. Some practical devices have been fabricated by MOF
crystals or thin films. For example, Chen et al. synthesized the twofold interpenetrating
framework [Zn2(bdc)2(bpy)] (MOF-508), in which the narrow channel (4.0  4.0 Å2)
is slightly larger than that of a methane molecule (3.8  3.8 Å2). A gas chro-
matographic (GC) column was packed with crystals of guest-free MOF-508, which
can efficiently separate linear and branched alkanes based on a molecular sieving
effect [71]. Using nanocrystals of MAF-4, Yan et al. fabricated capillary GC columns,
which showed high-resolution chromatographic separation of linear alkanes [117].
The high porosity and structural regularity of MOFs are suitable for molecular
enrichment and separation. A few MOFs have been used for solid-phase
microextraction (SPME), in which porous organic polymers are traditionally
employed to coat the extracting fiber. The first MOF-coated SPME fiber was
reported by Yan et al. using HKUST-1, which showed large enrichment and low
limits of detection for benzene homologues, but the extraction efficiency is largely
reduced at humidity over 30% due to the poor water stability of HKUST-1 [118].
They also fabricated highly water-stable SPME fibers based on MAF-4 and MAF-3.
Due to the very small aperture sizes of MAF-4 (3.4 Å) and MAF-3 (2.9 Å), the
former showed superior selectivity for n-alkanes over branched alkanes, while
the latter adsorbed neither, and larger molecules such as benzene homologues can
be only adsorbed on the outer surfaces [119]. Using 4-(3,5-dimethylpyrazol-4-yl)
benzoic acid (H2mpba) as a bifunctional ligand, Zhang et al. synthesized a highly
stable and porous MOF [Zn(mpba)] (MAF-X8), which possesses large and regular
1D channels with a relatively hydrophobic pore surface. MAF-X8 can be directly
grown on the surface of stainless-steel fiber to form thin layers with channels
exposed, which showed not only extraction efficiency for benzene homologues
much higher than conventional organic polymer-coated SPME fibers but also
superior selectivity even in the presence of large amounts of phenols. As the
benzene homologues and phenols have similar molecular sizes, the molecular
Metal–Organic Frameworks: From Design to Materials 21

sieving effect was attributed to the difference of diffusion speed, which was proven
by time-dependent extraction experiments and molecular dynamic calculations
[120].

4.4 Molecular Sensing

The unique coordination structures of MOFs consisting of various metal ions/


clusters and organic ligands give rise to not only unique sorption behaviors but
also rich optic, electric, and magnetic properties [121, 122]. Frequently, these
physical properties are also influenced by the guest species or host–guest
interactions. Further, guest adsorption, desorption, and exchange can change the
host framework structures of flexible MOFs, which alter the physical properties
more significantly. In turn, the species and concentration of guest molecules can be
deduced by the change of structures and physical properties of MOF crystals
[123]. For example, Chen et al. showed that a microporous MOF
[KCo7(OH)3(ipa)6(H2O)3] (H2ipa ¼ isophthalic acid) can reversibly release/reload
the water molecule coordinated with the Co(II) ions in a single-crystal to single-
crystal transformation manner, during which some of the six-coordinated Co(II)
ions are changed into pseudo five- and four-coordinated ones. The color of the
crystals is reversibly changed between red and purple, being similar to that of
commonly used silica gels. The change of coordination geometry of the Co(II) ions
also alters the magnetic properties between superparamagnetism and paramagetism
[124]. Also, spin crossover, which is based on the high and low spin transition of
single metal ions and very sensitive to the coordination geometry, electronic
structure of organic ligand, and guest perturbation of crystal lattice, can be used
as mechanism for guest sensing [125–127].
Compared with other physical properties, the change of color or photolumi-
nescence can be more easily identified and measured. Color change of MOFs can be
originated from the intrinsic properties of the chromophore involved [128], or from
the crystallite/film structures of the devices [129, 130]. The rich photophysical
properties of metal complexes in both solution and solid states have been widely
studied for a long time. The luminophors are regularly immobilized in the crystal
lattices of MOFs with high concentration. The highly porous structures of MOFs
can enhance the interaction between the luminophors and guest molecules in the
confined space. A variety of guest species can be detected by photoluminescent
MOFs [122, 131]. For example, Chen et al. showed that small molecules such as
dimethyl formamide, acetone, and ethanol exhibit different enhancing and
quenching effects on the photoluminescence of a microporous MOF [Eu(btc)] by
binding to the Eu(III) ions [132]. Li et al. reported microporous MOFs with high
quenching efficiencies for organic compounds containing nitro groups, which can
be used to detect chemical explosives [133]. Lin et al. showed that the classic
phosphorescent precious metal-based complexes can be decorated as metallo-
ligands to construct MOFs highly sensitive to molecular oxygen [134].
22 J.-P. Zhang and X.-M. Chen

The interaction between luminophors can significantly change the luminescence


properties such as emission wavelengths and lifetimes. In solution, the luminophor–
luminophor interactions are generally controlled by the concentration. In MOFs, the
average separation between adjacent luminophors are much shorter than those
in solution, but the self-quenching effect or excimer formation can be avoided
because the luminophors are fixed by the coordination network. Nevertheless, the
luminophor–luminophor interaction can be rationally changed by framework
flexibility or doping. For example, 1H-imidazo[4,5-f][1, 10]phenanthroline (Hip)
is a large π-conjugated molecule displaying strong blue photoluminescence in the
discrete state (such as in dilute solution), but it is non-luminescent in the solid state
due to extensive intermolecular interactions and energy transfer. Zhang et al.
reported a 3D flexible porous framework [Zn7(ip)12](OH)2, in which the
ligand–ligand interaction can be gradually changed by guest-induced framework
distortion, giving large change of emission color from blue to orange [135].
MOFs also showed other types of photonic applications, which are discussed in
other chapters of this book. Besides the interesting properties and applications
discussed above, MOFs can serve as catalysts and catalyst supporters [136–138],
as reaction containers for small organic molecules and polymers [139, 140], as well
as biomedicine [141] and second-order nonlinear optics [142].

5 Conclusions

With the development in molecular design, such as the zeolitic and metal cluster-
based net strategies, and synthetic approaches in the last two decades, a large
number of very interesting MOFs have been uncovered. The considerable efforts
have demonstrated that MOFs have not only designable and diverse framework
structures with different pore sizes and shapes but also modifiable and flexible
frameworks. Such characteristics provide better molecular recognition abilities
associated with the unique framework flexibilities and host–guest molecular
interactions, hence exhibit higher performance than other porous materials. Mean-
while, the recent progress in fabrication of this kind of materials into devices
significantly shortens the road to practical applications.
We can anticipate that, by further efforts and, in particularly collaborating with
engineers, researchers will be able to design and prepare more practically applicable
MOF-based materials with unique pore surface, sizes, shapes, and framework
flexibilities for excellent molecular recognition abilities and higher performance
than other porous materials, in adsorption, separation, sensing, catalysis and other
aspects, and to finely modify and fabricate these materials into practical devices.

Acknowledgment This work was supported by the “973 Project” (2012CB821706) and NSFC
(21290173, 21225105 and 21121061), Science and Technology Department of Guangdong Prov-
ince (S2012030006240).
Metal–Organic Frameworks: From Design to Materials 23

References

1. Hoskins BF, Robson R (1989) J Am Chem Soc 111:5962


2. Hoskins BF, Robson R (1990) J Am Chem Soc 112(1546):1554
3. Kondo M, Okubo T, Asami A, Noro S, Yoshitomi T, Kitagawa S, Ishii T, Matsuzaka H,
Seki K (1999) Angew Chem Int Ed 38:140
4. Li H, Eddaoudi M, O’Keeffe M, Yaghi OM (1999) Nature 402:276
5. Tranchemontagne DJ, Mendoza-Cortes JL, O’Keeffe M, Yaghi OM (2009) Chem Soc Rev
38:1257
6. Batten SR, Champness NR, Chen XM, Garcia-Martinez J, Kitagawa S, Ohrstrom L,
O’Keeffe M, Suh MP, Reedijk J (2012) CrystEngComm 14:3001
7. Ferey G (2008) Chem Soc Rev 37:191
8. Czaja AU, Trukhan N, Muller U (2009) Chem Soc Rev 38:1284
9. Horike S, Shimomura S, Kitagawa S (2009) Nat Chem 1:695
10. Robson R (2000) J Chem Soc Dalton Trans 3735
11. Yaghi OM, O’Keeffe M, Ockwig NW, Chae HK, Eddaoudi M, Kim J (2003) Nature 423:705
12. Wells AF (1977) Three-dimensional nets and polyhedra. Wiley, New York
13. Reticular chemistry structure resource (RCSR). http://rcsr.anu.edu.au/
14. O’Keeffe M, Peskov MA, Ramsden SJ, Yaghi OM (2008) Acc Chem Res 41:1782
15. Blatov VA, O’Keeffe M, Proserpio DM (2010) CrystEngComm 12:44
16. IZA structure commission. http://izasc.ethz.ch/fmi/xsl/IZA-SC/ft.xsl
17. Zhang J-P, Huang X-C, Chen X-M (2009) Chem Soc Rev 38:2385
18. Zhang JP, Zhang YB, Lin JB, Chen XM (2012) Chem Rev 112:1001
19. Biradha K, Hongo Y, Fujita M (2002) Angew Chem Int Ed 41:3395
20. Fletcher AJ, Cussen EJ, Bradshaw D, Rosseinsky MJ, Thomas KM (2004) J Am Chem Soc
126:9750
21. Noro S, Kitagawa S, Kondo M, Seki K (2000) Angew Chem Int Ed 39:2082
22. Evans OR, Lin WB (2002) Acc Chem Res 35:511
23. Eddaoudi M, Kim J, O’Keeffe M, Yaghi OM (2002) J Am Chem Soc 124:376
24. Zhang J-P, Zheng S-L, Huang X-C, Chen X-M (2004) Angew Chem Int Ed 43:206
25. Zhang JP, Lin YY, Huang XC, Chen XM (2005) J Am Chem Soc 127:5495
26. Zhang J-P, Chen X-M (2008) J Am Chem Soc 130:6010
27. Batten SR, Hoskins BF, Moubaraki B, Murray KS, Robson R (2000) Chem Commun 1095
28. Dybtsev DN, Chun H, Kim K (2004) Angew Chem Int Ed 43:5033
29. Zheng YZ, Tong ML, Xue W, Zhang WX, Chen XM, Grandjean F, Long GJ (2007)
Angew Chem Int Ed 46:6076
30. Zhang YB, Zhang WX, Feng FY, Zhang JP, Chen XM (2009) Angew Chem Int Ed 48:5287
31. Zhang YB, Zhou HL, Lin RB, Zhang C, Lin JB, Zhang JP, Chen XM (2012) Nat Comm 3:642
32. Chui SSY, Lo SMF, Charmant JPH, Orpen AG, Williams ID (1999) Science 283:1148
33. Chen BL, Eddaoudi M, Hyde ST, O’Keeffe M, Yaghi OM (2001) Science 291:1021
34. Wang XS, Ma SQ, Yuan DQ, Yoon JW, Hwang YK, Chang JS, Wang XP, Jorgensen MR,
Chen YS, Zhou HC (2009) Inorg Chem 48:7519
35. Eddaoudi M, Kim J, Rosi N, Vodak D, Wachter J, O’Keeffe M, Yaghi OM (2002) Science
295:469
36. Hou L, Lin YY, Chen XM (2008) Inorg Chem 47:1346
37. Feréy G, Mellot-Draznieks C, Serre C, Millange F, Dutour J, Surble S, Margiolaki I (2005)
Science 309:2040
38. Serre C, Mellot-Draznieks C, Surble S, Audebrand N, Filinchuk Y, Ferey G (2007)
Science 315:1828
39. Perry JJ, Perman JA, Zaworotko MJ (2009) Chem Soc Rev 38:1400
40. Eddaoudi M, Kim J, Wachter JB, Chae HK, O’Keeffe M, Yaghi OM (2001) J Am Chem Soc
123:4368
41. Ni Z, Yassar A, Antoun T, Yaghi OM (2005) J Am Chem Soc 127:12752
24 J.-P. Zhang and X.-M. Chen

42. Chun H (2008) J Am Chem Soc 130:800


43. Chun H, Jung HJ, Seo JW (2009) Inorg Chem 48:2043
44. Li JR, Timmons DJ, Zhou HC (2009) J Am Chem Soc 131:6368
45. Zhang XM, Zheng YZ, Li CR, Zhang WX, Chen XM (2007) Cryst Growth Des 7:980
46. Jia JH, Lin X, Wilson C, Blake AJ, Champness NR, Hubberstey P, Walker G, Cussen EJ,
Schroder M (2007) Chem Commun 840
47. Wei Y-S, Chen K-J, Liao P-Q, Zhu B-Y, Lin R-B, Zhou H-L, Wang B-Y, Xue W, Zhang J-P,
Chen X-M (2013) Chem Sci 4:1539
48. Sun DF, Ma SQ, Ke YX, Collins DJ, Zhou HC (2006) J Am Chem Soc 128:3896
49. Ma SQ, Sun DF, Ambrogio M, Fillinger JA, Parkin S, Zhou HC (2007) J Am Chem Soc
129:1858
50. Ma SQ, Eckert J, Forster PM, Yoon JW, Hwang YK, Chang JS, Collier CD, Parise JB,
Zhou HC (2008) J Am Chem Soc 130:15896
51. Takashima Y, Martinez VM, Furukawa S, Kondo M, Shimomura S, Uehara H, Nakahama M,
Sugimoto K, Kitagawa S (2011) Nat Comm 2:168
52. Chen BL, Fronczek FR, Maverick AW (2003) Chem Commun 2166
53. Bu XH, Tong ML, Chang HC, Kitagawa S, Batten SR (2004) Angew Chem Int Ed 43:192
54. Lin J-B, Zhang J-P, Chen X-M (2010) J Am Chem Soc 132:6654
55. Sturm M, Brandl F, Engel D, Hoppe W (1975) Acta Crystallogr B 31:2369
56. Lehnert R, Seel F (1980) Z Anorg Allg Chem 464:187
57. Tian YQ, Cai CX, Ji Y, You XZ, Peng SM, Lee GH (2002) Angew Chem Int Ed 41:1384
58. Tian YQ, Cai CX, Ren XM, Duan CY, Xu Y, Gao S, You XZ (2003) Chem Eur J 9:5673
59. Huang XC, Zhang JP, Chen XM (2004) J Am Chem Soc 126:13218
60. Huang X-C, Zhang J-P, Lin Y-Y, Chen X-M (2005) Chem Commun 2232
61. Huang X-C, Zhang J-P, Chen X-M (2006) Cryst Growth Des 6:1194
62. Zhang JP, Qi XL, He CT, Wang Y, Chen XM (2011) Chem Commun 47:4156
63. Huang X-C, Zhang J-P, Chen X-M (2003) Chin Sci Bull 48:1531
64. Li YS, Liang FY, Bux H, Feldhoff A, Yang WS, Caro J (2010) Angew Chem Int Ed 49:548
65. Huang X-C (2004) PhD Thesis thesis, Sun Yat-Sen University, Guangzhou
66. Huang XC, Lin YY, Zhang JP, Chen XM (2006) Angew Chem Int Ed 45:1557
67. Banerjee R, Phan A, Wang B, Knobler C, Furukawa H, O’Keeffe M, Yaghi OM (2008)
Science 319:939
68. Phan A, Doonan CJ, Uribe-Romo FJ, Knobler CB, O’Keeffe M, Yaghi OM (2010) Acc Chem
Res 43:58
69. Zhang J-P, Zhu A-X, Lin R-B, Qi X-L, Chen X-M (2011) Adv Mater 23:1268
70. Zhang J-P, Chen X-M (2006) Chem Commun 1689
71. Chen BL, Liang CD, Yang J, Contreras DS, Clancy YL, Lobkovsky EB, Yaghi OM, Dai S
(2006) Angew Chem Int Ed 45:1390
72. Lee JY, Pan L, Huang XY, Emge TJ, Li J (2011) Adv Funct Mater 21:993
73. Wang XF, Wang Y, Zhang YB, Xue W, Zhang JP, Chen XM (2012) Chem Commun 48:133
74. Stock N, Biswas S (2012) Chem Rev 112:933
75. Fukushima T, Horike S, Inubushi Y, Nakagawa K, Kubota Y, Takata M, Kitagawa S (2010)
Angew Chem Int Ed 49:4820
76. Deng HX, Doonan CJ, Furukawa H, Ferreira RB, Towne J, Knobler CB, Wang B, Yaghi OM
(2010) Science 327:846
77. Jhung SH, Lee JH, Forster PM, Ferey G, Cheetham AK, Chang JS (2006) Chem Eur J
12:7899
78. Kole GK, Vittal JJ (2013) Chem Soc Rev 42:1755
79. Lin J-B, Lin R-B, Cheng X-N, Zhang J-P, Chen X-M (2011) Chem Commun 47:9185
80. Zhu AX, Lin RB, Qi XL, Liu Y, Lin YY, Zhang JP, Chen XM (2012) Microporous
Mesoporous Mat 157:42
81. Shimomura S, Matsuda R, Tsujino T, Kawamura T, Kitagawa S (2006) J Am Chem Soc
128:16416
Metal–Organic Frameworks: From Design to Materials 25

82. Shimomura S, Horike S, Matsuda R, Kitagawa S (2007) J Am Chem Soc 129:10990


83. Chen XM, Tong ML (2007) Acc Chem Res 40:162
84. Zhang J-P, Chen X-M (2009) In: Hong M-C, Chen L (eds) Design and construction of
coordination polymers, 86th edn. Wiley, New Jersey, p 63
85. Zheng YZ, Xue W, Tong ML, Chen XM, Grandjean F, Long GJ (2008) Inorg Chem 47:4077
86. Zheng YZ, Xue W, Zhang WX, Tong ML, Chen XM, Grandjean F, Long GJ, Ng SW,
Panissod P, Drillon M (2009) Inorg Chem 48:2028
87. Wang YT, Fan HH, Wang HZ, Chen XM (2005) Inorg Chem 44:4148
88. Wang ZQ, Cohen SM (2009) Chem Soc Rev 38:1315
89. Cohen SM (2012) Chem Rev 112:970
90. Kitagawa S, Kitaura R, Noro S (2004) Angew Chem Int Ed 43:2334
91. Kitagawa S, Uemura K (2005) Chem Soc Rev 34:109
92. Ferey G, Serre C (2009) Chem Soc Rev 38:1380
93. Gould SL, Tranchemontagne D, Yaghi OM, Garcia-Garibay MA (2008) J Am Chem Soc
130:3246
94. He J, Yin YG, Wu T, Li D, Huang XC (2006) Chem Commun 2845
95. Zhang J-P, Horike S, Kitagawa S (2007) Angew Chem Int Ed 46:889
96. Zhang JP, Kitagawa S (2008) J Am Chem Soc 130:907
97. Xie LH, Lin JB, Liu XM, Wang Y, Zhang WX, Zhang JP, Chen XM (2010) Inorg Chem
49:1158
98. Xie LH, Wang Y, Liu XM, Lin JB, Zhang JP, Chen XM (2011) CrystEngComm 13:5849
99. Zhang JP, Zhu AX, Chen XM (2012) Chem Commun 48:11395
100. Sakata Y, Furukawa S, Kondo M, Hirai K, Horike N, Takashima Y, Uehara H, Louvain N,
Meilikhov M, Tsuruoka T, Isoda S, Kosaka W, Sakata O, Kitagawa S (2013) Science 339:193
101. Murray LJ, Dincă M, Long JR (2009) Chem Soc Rev 38:1294
102. Suh MP, Park HJ, Prasad TK, Lim DW (2012) Chem Rev 112:782
103. Ma SQ, Sun DF, Simmons JM, Collier CD, Yuan DQ, Zhou HC (2008) J Am Chem Soc
130:1012
104. Bhatia SK, Myers AL (2006) Langmuir 22:1688–1700
105. Furukawa H, Miller MA, Yaghi OM (2007) J Mater Chem 17:3197
106. Matsuda R, Kitaura R, Kitagawa S, Kubota Y, Belosludov RV, Kobayashi TC, Sakamoto H,
Chiba T, Takata M, Kawazoe Y, Mita Y (2005) Nature 436:238
107. Zhang J-P, Chen X-M (2009) J Am Chem Soc 131:5516
108. Li J-R, Kuppler RJ, Zhou H-C (2009) Chem Soc Rev 38:1477
109. Li JR, Sculley J, Zhou HC (2012) Chem Rev 112:869
110. Wu HH, Gong QH, Olson DH, Li J (2012) Chem Rev 112:836
111. Guo HL, Zhu GS, Hewitt IJ, Qiu SL (2009) J Am Chem Soc 131:1646
112. Venna SR, Carreon MA (2010) J Am Chem Soc 132:76
113. Sumida K, Rogow DL, Mason JA, McDonald TM, Bloch ED, Herm ZR, Bae TH, Long JR
(2012) Chem Rev 112:724
114. Liao PQ, Zhou DD, Zhu AX, Jiang L, Lin RB, Zhang JP, Chen XM (2012) J Am Chem Soc
134:17380
115. Wriedt M, Sculley JP, Yakovenko AA, Ma YG, Halder GJ, Balbuena PB, Zhou HC (2012)
Angew Chem Int Ed 51:9804
116. Li K, Olson DH, Seidel J, Emge TJ, Gong H, Zeng H, Li J (2009) J Am Chem Soc 131:10368
117. Chang N, Gu ZY, Yan XP (2010) J Am Chem Soc 132:13645
118. Cui XY, Gu ZY, Jiang DQ, Li Y, Wang HF, Yan XP (2009) Anal Chem 81:9771
119. Chang N, Gu ZY, Wang HF, Yan XP (2011) Anal Chem 83:7094
120. He CT, Tian JY, Liu SY, Ouyang GF, Zhang JP, Chen XM (2013) Chem Sci 4:351
121. Kurmoo M (2009) Chem Soc Rev 38:1353
122. Allendorf MD, Bauer CA, Bhakta RK, Houk RJT (2009) Chem Soc Rev 38:1330
123. Kreno LE, Leong K, Farha OK, Allendorf M, Van Duyne RP, Hupp JT (2012) Chem Rev
112:1105
26 J.-P. Zhang and X.-M. Chen

124. Cheng XN, Zhang WX, Lin YY, Zheng YZ, Chen XM (2007) Adv Mater 19:1494
125. Ohba M, Yoneda K, Agusti G, Munoz MC, Gaspar AB, Real JA, Yamasaki M, Ando H,
Nakao Y, Sakaki S, Kitagawa S (2009) Angew Chem Int Ed 48:4767
126. Lin JB, Xue W, Wang BY, Tao J, Zhang WX, Zhang JP, Chen XM (2012) Inorg Chem
51:9423
127. Bao X, Shepherd HJ, Salmon L, Molnar G, Tong ML, Bousseksou A (2013) Angew Chem Int
Ed 52:1198
128. Lu ZZ, Zhang R, Li YZ, Guo ZJ, Zheng HG (2011) J Am Chem Soc 133:4172
129. Lu G, Hupp JT (2010) J Am Chem Soc 132:7832
130. Lu G, Farha OK, Zhang WN, Huo FW, Hupp JT (2012) Adv Mater 24:3970
131. Cui YJ, Yue YF, Qian GD, Chen BL (2012) Chem Rev 112:1126
132. Chen BL, Yang Y, Zapata F, Lin GN, Qian GD, Lobkovsky EB (2007) Adv Mater 19:1693
133. Lan AJ, Li KH, Wu HH, Olson DH, Emge TJ, Ki W, Hong MC, Li J (2009) Angew Chem Int
Ed 48:2334
134. Xie ZG, Ma LQ, deKrafft KE, Jin A, Lin WB (2010) J Am Chem Soc 132:922
135. Qi XL, Lin RB, Chen Q, Lin JB, Zhang JP, Chen XM (2011) Chem Sci 2:2214
136. Ma LQ, Abney C, Lin WB (2009) Chem Soc Rev 38:1248
137. Lee J, Farha OK, Roberts J, Scheidt KA, Nguyen ST, Hupp JT (2009) Chem Soc Rev 38:1450
138. Yoon M, Srirambalaji R, Kim K (2012) Chem Rev 112:1196
139. Kawamichi T, Kodama T, Kawano M, Fujita M (2008) Angew Chem Int Ed 47:8030
140. Uemura T, Yanai N, Kitagawa S (2009) Chem Soc Rev 38:1228
141. Horcajada P, Gref R, Baati T, Allan PK, Maurin G, Couvreur P, Ferey G, Morris RE, Serre C
(2012) Chem Rev 112:1232
142. Wang C, Zhang T, Lin WB (2012) Chem Rev 112:1084
Struct Bond (2014) 157: 27–88
DOI: 10.1007/430_2013_133
# Springer-Verlag Berlin Heidelberg 2013
Published online: 7 September 2013

Luminescent Properties and Applications


of Metal-Organic Frameworks

Yuanjing Cui, Banglin Chen, and Guodong Qian

Abstract Metal-organic frameworks (MOFs) are very promising multifunctional


luminescent materials because of their inherent advantages of both organic linkers
and inorganic metal ions, as well as the tailorability in terms of structure, dimension,
size, and shape. Furthermore, the metal centers, organic linkers, metal-organic charge
transfer, and guest molecules within porous MOFs all can potentially generate
luminescence. Such uniqueness can allow us to generate luminescent MOF materials
with systematically varied luminescence properties which are crucial for the lighting,
display, and optical devices. Although luminescent MOFs are still in their infancy,
the currently available results have unambiguously demonstrated that the design and
construction of MOFs for luminescent functionality is very active, and hundreds of
papers published in the last few years indicate that this interest is still increasing. This
chapter outlines the basic principles and luminescence origins of MOF materials and
then describes the latest developments in luminescent MOFs and their applications in
light-emitting devices, chemical sensors, and biomedicine. As a conclusion, several
evident research trends were proposed which are particularly promising but will
require concerted effort for success.

Keywords Biomedicine  Chemical sensor  Luminescence  Metal-organic


frameworks  White-light-emitting

Y. Cui and G. Qian (*)


State Key Laboratory of Silicon Materials, Department of Materials Science and Engineering,
Zhejiang University, Hangzhou 310027, China
e-mail: gdqian@zju.edu.cn
B. Chen (*)
State Key Laboratory of Silicon Materials, Department of Materials Science and Engineering,
Zhejiang University, Hangzhou 310027, China
Department of Chemistry, University of Texas at San Antonio, San Antonio, TX 78249-0698,
USA
e-mail: Banglin.Chen@utsa.edu
28 Y. Cui et al.

Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2 Construction and Modification of MOFs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.1 Syntheses of MOFs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.2 Nanoscale MOFs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.3 Postsynthesis Modification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3 Luminescent Properties of MOFs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.1 Ligand-Based Luminescence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.2 Lanthanide Luminescence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.3 Charge-Transfer Luminescence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.4 Guest-Induced Luminescence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4 Applications of Luminescent MOF . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.1 Near-Infrared (NIR) Luminescent Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.2 White-Light-Emitting Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.3 Chemical Sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.4 Application in Biomedicine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
5 Conclusions and Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

Abbreviations

1,2-BDC 1,2-Benzenedicarboxylate
1,3-BDC 1,3-Benzenedicarboxylate
1,4-BDC 1,4-Benzenedicarboxylate
1,4-NDC 1,4-Naphthalenedicarboxylate
2,3-pydcH2 Pyridine-2,3-dicarboxylic acid
2,4-DNT 2,4-Dinitrotoluene
2,6-pydc Pyridine-2,6-dicarboxylate
3-Abpt 4-Amine-3,5-bis(3-pyridyl)-1,2,4-triazole
3-PyHBIm 2-(3-Pyridyl)benzimidazole
4-Ptz 5-(4-Pyridyl)tetrazole
4-PyHBIm 2-(4-Pyridyl)benzimidazole
5MT 5-Methyl-tetrazole
ABTC 3,30 ,5,50 -Azobenzenetetracarboxylate
ATA 5-Amino-tetrazole
ATPA 2-Aminoterephthalate
BCPA 9,10-Bis( p-carboxyphenyl) anthracene
BDC-F4 2,3,5,6-Tetrafluoro-1,4-benzenedicarboxylate
bhc Benzenehexacarboxylate
bipy 4,40 -Bipyridine
bpdc 4,40 -Biphenyldicarboxylate
Bpe Trans-bis(4-pyridyl)ethylene
bpee 1,2-Bipyridylethene
bpp 1,3-Bis(4-pyridyl)propane
BPT Biphenyl-3, 40 , 5-tricarboxylate
bpydc 2,20 -Bipyridine-5,50 -dicarboxylate
Luminescent Properties and Applications of Metal-Organic Frameworks 29

bpz 3,30 ,5,50 -Tetramethyl-4,40 -bipyrazolate


bpze Bis(pyrazole-1-yl)ethane
bpzm Bis(pyrazole-1-yl)methane
bpzp Bis(pyrazole-1-yl)prothane
BTAH Benzotriazole
BTB 4,40 ,400 -Benzenetribenzoate
BTC 1,3,5-Benzenetricarboxylate
CCT Correlated color temperature
chdc 1,4-Cyclohexanedicarboxylate
CIE Commission International ed’Eclairage
CRI Color rendering index
CT Computed tomography
CTAB Cetyltrimethylammonium bromide
CTC Cis,cis-1,3,5-cyclohexanetricarboxylate
DABCO 1,4-Diazabicyclo[2.2.2]octane
dbm Dibenzoylmethane
DFDA 9,9-Dipropylfluorene-2,7-dicarboxylate
DHT 2,5-Dihydroxyterephthalate
dipicH2 Dipicolinic acid
DMA Dimethylammonium
DMBDC 2,5-Dimethoxy-1,4-benzenedicarboxylate
DMNB 2,3-Dimethyl-2,3-dinitrobutane
dpbp 4,40 -Bis(diphenylphosphoryl) biphenyl
dpNDI N,N0 -di(4-pyridyl)-1,4,5,8-naphthalenediimide
ESCP Ethoxysuccinato-cisplatin
FMA Fumarate
H2dtba 2,20 -Dithiobisbenzoic acid
H2ox Oxalic acid
H2pda 1,4-Phenylenediacetic acid
H2PhenDCA 1,10-Phenanthroline-2,9-dicarboxylic acid
H2-PVDC 4,40 -[(2,5-Dimethoxy-1,4-phenylene)di-2,1-ethenediyl]bis-benzoic
acid
H3bidc 1H-benzimidazole-5,6-dicarboxylic acid
H3NTB 4,40 ,400 -Nitrilotrisbenzoic acid
hfa Hexafluoro acetylacetonato
HMOFs Heterometal-organic frameworks
Hnic Nicotinic acid
IDC Imidazole-4,5-dicarboxylate
ILAG Ion- and liquid-assisted grinding
imdc 4,5-Imidazoledicarboxylate
LAG Liquid-assisted grinding
LMCT Ligand-to-metal charge transfer
Ln3+ Lanthanide ions
MES 2-(N-morpholino)ethanesulfonate
30 Y. Cui et al.

MLCT Metal-to-ligand charge transfer


MOFs Metal-organic frameworks
MRI Magnetic resonance imaging
nds 2,6-Naphthanlenedisulfonate
NIR Near-infrared
NMOFs Nano-MOFs
NMP 1-Methyl-2-pyrrolidone
NTA Nitrilotriacetate
OBA 4,40 -Oxybis(benzoate)
ox Oxalate
PDA Pyridine-2,6-dicarboxylic acid
pdc 2,5-Pyridinedicarboxylate
Phen 1,10-Phenanthroline
PhPPy2 Bis(2-pyridyl)phenylphosphine
pimda 2-Propyl-4,5-imidazole-dicarboxylate
Pmtz 5-(Pyrimidyl)-tetrazolato
ppy 2-Phenylpyridine
pta 2,4,6-Pyridinetricarboxylate
pydc Pyridine-dicarboxylate
Pz 3,5-Dimethylpyrazole
pzdc 2,5-Pyrazinedicarboxylate
QD Quantum dot
quin 2,20 -Biquinoline
Rh6G Rhodamine 6G
SBUs Secondary building units
sfdb 4,40 -Sulfonyldibenzoic acid
TAA 1-H-tetrazole-5-acetate
TBAPy 1,3,6,8-Tetrakis( p-benzoic acid)pyrene
TDC Thiophene-2,5-dicarboxylate
TEAF Triethylammonium fluoride
TNT 2,4,6-Trinitrotoluene
tta 2-Thenoyltrifluoroacetone
TzC 5-Carboxylato-tetrazolato
UV Ultraviolet
VOCs Volatile organic compounds

1 Introduction

Metal-organic frameworks (MOFs), also known as coordination polymers or coordi-


nation networks, have attracted tremendous attention over the last decade due to their
promising applications in gas storage [1, 2], gas capture [3], separations [4], chemical
catalysis [5], luminescence [6–8], magnetism [9], and drug delivery [10]. MOFs are
highly crystalline inorganic–organic hybrid materials comprised of metal ions or
Luminescent Properties and Applications of Metal-Organic Frameworks 31

clusters that form vertices of a framework and organic linkers that act as bridges. In
most cases, MOFs are rigid enough to allow the removal of solvent or other guest
molecules in the pores resulting in permanent porosity. Thus, these materials generally
exhibit microporous character, and the pore sizes could be tuned from several
angstroms to several nanometers by altering the length of organic linkers or controlling
the interpenetration. The extraordinarily high porosity, very large surface areas, and
wide range of pore sizes of MOFs make them attractive for scientific and industrial
application as new porous materials [11–15]. Compared with conventional porous
materials, such as zeolites, carbon-based materials, one main advantage of MOFs is
their easily modifiable synthesis to control the size, shape, dimension, connectivity,
and surface of the pores by changing the organic ligands and metal ions. A variety of
functionality can be incorporated in the MOFs by varying the coordination geometries
of the metals and the topicities of the ligands, modifying the pore surface, or
introducing functional groups. These features are attractive enough to attract more
research effort toward the development of MOFs with designable framework
structures and tunable functionality for specific applications.
Although some early reports on MOFs have appeared in 1960s, the interest in
this field was kindled when Yaghi et al. reported the structure of the well-known
MOF-5 with significantly high surface area of 2,900 m2/g in 1999 [16]. The
implementation of simple solvothermal synthesis and incorporation of large
organic linkers to assemble highly porous MOF-14 in 2001 [17], to some extent,
has highlighted the establishment of a general strategy to construct porous MOFs.
Since then, much research effort has been devoted for the investigation of the high
porosity, stability, control of pore shape and size, and framework flexibility. Up to
date, thousands of different MOFs have been synthesized, and the number of
publications related with MOFs has increased remarkably.
Recently, the development of luminescent functional MOFs has attracted more
scientific interest because both the inorganic and organic moieties can provide the
platforms to generate luminescence, while metal–ligand charge-transfer-related
luminescence within MOFs can add another dimensional luminescent functionality.
Furthermore, some guest molecules within MOFs can also emit and/or induce
luminescence [7, 18–20]. The infinite three-dimensional (3D) framework
structures, inherent microporosity, and high surface areas offer great opportunities
for intermolecular interaction between MOFs and guest molecules, which can
dramatically vary the coordination environment and energy transfer, thus
modulating the luminescent properties of MOFs. Moreover, the pore surface and
framework structure of the MOFs can be further functionalized by postsynthetic
modification, which enables the energy level and charge transfer to be finely tuned
targeting a specific luminescent property. Compared with the traditional molecular
approach, by making use of molecular organic and coordination compounds to
explore functional luminescent materials, the MOF approach also has the advantage
to collaboratively interplay/interact with each other among the periodically located
luminescent centers to direct new functionalities. These advantages lead to the vast
versatility of luminescent MOFs in practical applications, which has exceeded the
capabilities of many other luminescent materials.
32 Y. Cui et al.

During the past 10 years, there is an increasing trend on the exploration and
discovery of functional luminescent MOFs. Studies on luminescent MOFs up to
now mainly include the following: investigating the fundamental synthesis and
luminescent properties of MOFs; creating MOFs with tunable luminescence for
application in light-emitting and display devices; using luminescence of MOFs to
probe local environment, structure, and guest species; and developing multifunctional
MOFs combining luminescence, magnetism, and biocompatibility for biomedical
application. The permanent porosity with some luminescent MOFs has enabled their
reversible storage and release of guest substrates and provided the hosts for their
differential recognitions with sensing species, thus distinguishing porous luminescent
MOFs from the traditional inorganic and organic luminescent materials in terms of
their multifunctionalities. The MOF approach can also offer a variety of other
attractive characteristics such as the straightforward syntheses, predictable structures
and porosities, nanoscale processibility, and collaborative properties to develop lumi-
nescent functional MOF materials. The chapter describes the latest developments in
luminescent MOF materials and their applications in light-emitting devices, chemical
sensors, and biomedicine. As a conclusion, several evident research trends were
proposed which are particularly promising but will require concerted effort for
success. The purpose of this chapter is to give a comprehensive overview of the
different types of luminescent MOFs, particularly those with interesting and important
applications. The definition of MOFs is interpreted rather broadly in this review, so
some nonporous coordination polymers will be also included.

2 Construction and Modification of MOFs

2.1 Syntheses of MOFs

MOFs are essentially constructed through a simple approach known as the building-
block methodology, which relies on the assembly of molecular or ionic moieties
into extended arrays using complementary interactions between components. In the
construction of MOFs, one-dimensional (1D) chain structure can be achieved
through the connecting of metal ions that prefer linear coordination with simple
organic linkers, while two-dimensional (2D) sheet structures and 3D architectures
can be built using trigonal or square-planar metal center nodes and tetrahedral or
octahedral metal centers with the same organic linkers, respectively. Thus, by
changing the metal geometry and the organic linker type, a wide variety of different
MOFs can be generated. In addition, the in situ formation of some metal-containing
clusters (generally termed as secondary building units (SBUs)) with predetermined
coordination geometries during the MOF syntheses has enabled us to synthesize a
series of MOFs with predictable structures. For example, the connectivity of Zn4O
(COO)6 SBUs with a large number of bicarboxylate led to the so-called isoreticular
MOFs (IRMOFs) with the default cubic structures [21]. Another well-known
example of isoreticular MOFs are those assembled from paddle-wheel M2(COO)4
Luminescent Properties and Applications of Metal-Organic Frameworks 33

(M¼Cu2+, Zn2+, Co2+, and Ni2+) with bicarboxylate (R(COO)2) and pillar organic
linker (L) to form distorted primitive cubic nets of M2(ROO)2L [22, 23].
The compound 4,40 -bipyridine or its derivatives and multidentate carboxylic
acids with suitable spacers, especially benzene-multicarboxylic acids, are the
prominent linkers for constructing diverse architectures. The 4,40 -bipyridine-like
ligands are simple linear linkers which possess monodentate donor nitrogen atoms
at opposite ends of the molecule; therefore the determination of framework is
directly governed by the coordination environment and geometry of the metal
ion. Multidentate carboxylic acids are also attractive linkers because their high
acidity allows facile in situ deprotonation. In addition, the ability of carboxylic
acids to form more than one coordination bond increased the chances of generating
a robust porous framework that retains the MOF integrity in the absence of guest
molecules. Yaghi et al. have designed and synthesized sixteen MOFs, denoted as
IRMOF-n (n ¼ 1–16), with wide ranges of pore size and chemical functionality
through the reaction of various benzenedicarboxylate ligands with Zn2+ ions. The
framework consists of oxo-centered Zn4O tetrahedron cluster located at the edges
of the cube and bridged by six carboxylate units to give secondary building units
(SBUs), which are further connected by the bridging dicarboxylate, resulting in the
3D cubic networks. The MOFs with bigger cavities or channels can be synthesized
by using longer benzenedicarboxylate linkers [21]. Besides the rigid linkers as
mentioned above, flexible ligands such as cyclohexane-1,2,5,6-tetracarboxylic
acid, tetrakis(imidazole), and N,N0 -diacetic acid imidazolium can be used in the
synthesis of MOFs [24–27]. These linkers are prone to give nonporous structures
through dense packing of a single network or through interpenetration of two or
more networks.
Generally, MOFs can be produced by mixing two solutions containing the metal
ions and the organic ligands, either at room temperature or under hydrothermal/
solvothermal conditions [28–30]. Initial syntheses of coordination polymers/metal-
organic frameworks were realized by the diffusion technique in which the solution
containing metal salts was slowly diffused into the solution containing organic
linkers such as 4,4-bipyridyl, and/or the weak bases such as triethylamine were
slowly diffused into the solution containing metal salt and bridging organic carbox-
ylic acid [16]. Such synthesis approach is very time-consuming while the yield is
typically low. In fact, the yield to synthesize MOF-5 in the first published paper is
only about 5% [16]. Hydrothermal syntheses of MOFs have been feasible to those
reaction systems in which organic linkers can be partially soluble in water at higher
temperature. The implementation of solvothermal syntheses by making use of
organic solvents to readily dissolve organic linkers, as exemplified in the synthesis
of MOF-14 in 2001, has facilitated the syntheses of crystalline MOF materials
significantly [17]. Generally speaking, heating the mixture containing inorganic
salts and organic linkers in certain types of solvents (DMF is the most popular
solvent) at the certain temperature (60–120 C) for several hours to two days can
readily lead to crystalline MOFs for structure characterization and property explo-
ration. The simple and straightforward syntheses of MOFs, to some of the extent,
have allowed many chemists and materials scientists readily to start their research
34 Y. Cui et al.

programs in such a very active research field [11, 20, 31–33]. The solvothermal
synthesis approach has also enabled the exploration of combinatorial approach to
screen and discover new MOF materials within a very short period of time, as
exemplified in the combinatorial syntheses of a series of zeolitic imidazole
frameworks (ZIFs, a special subclass of MOFs) [34–39]. BASF has even been
able to produce and commercialize some prototypical MOFs in large scale.
In recent years, other alternative synthetic methods including mechanochemical
methods of solid-state grinding and liquid-assisted grinding, sonochemical methods,
and microwave-assisted synthesis have also been emerged. In mechanochemical
synthesis, a mixture of organic linkers and metal salts is ground together in a mechani-
cal ball mill to yield the MOFs [40–49]. Compared with solvothermal synthesis, the
obvious advantage of mechanochemical synthesis is the ability to avoid the use of
solvent. In addition, sufficient amounts of pure MOFs for broad-range testing can be
easily produced. Klimakow et al. demonstrated the mechanochemical synthesis of
the intensively studied MOFs HKUST-1 (Cu3(BTC)2, BTC ¼ 1,3,5-benzenetri-
carboxylate) and MOF-14 (Cu3(BTB)2, BTB ¼ 4,40 ,400 -benzenetribenzoate)
[41]. The obtained MOFs exhibit the high surface area which is comparable to the
highest given values in the literature for the respective compounds. Friščić
et al. introduced an improved mechanochemical approach, denoted as ion- and
liquid-assisted grinding (ILAG) [44]. The construction of MOFs is accelerated and
directed by small amounts of salts, for example, nitrate and sulfates induce the
formation of tetragonal and hexagonal structure, respectively.
Sonochemistry was recently explored by Son et al. in the synthesis of the popular
MOF-5 from solutions of zinc nitrate and terephthalic acid [50]. Compared with
conventional solution synthesis, sonochemistry method can lead to homogeneous
nucleation and a significant reduction in crystallization time [51–54]. By using
1-methyl-2-pyrrolidone (NMP) as solvent, high-quality MOF-5 crystals of 5–25 μm
in size are produced, with similar physicochemical properties to those crystals
prepared under conventional convective heating. Similarly, high-quality
MOF-177 crystals, namely, Zn4O(BTB)2 (BTB ¼ 4,40 ,400 -benzenetribenzoate),
are also synthesized via a sonochemical route in the presence of NMP as a solvent
[55]. The synthesis time is shortened drastically from 48 h to ca. 0.5 h accompanied
by a substantial reduction in crystal size to 5–20 μm. The product yield of MOF-177
synthesized via the sonochemical route was 95.6%, which was significantly higher
than the product yields of other methods.
Microwave-assisted solvothermal synthesis involves heating a solution with
microwaves to produce nanosized crystals, which allows high-quality MOF crystals
to be synthesized in under a minute [56–65]. Although the microwave method
usually cannot produce single crystals with a big size for structural analysis, its
homogeneous effects could create a uniform seeding condition; thus the size and
shape of the crystals can be well controlled and the rate of chemical reactions can be
accelerated. The application of microwave radiation to conduct MOF synthesis has
been investigated by Ni et al. in the construction of three known MOFs, namely,
IRMOF1 (MOF-5), IRMOF2, and IRMOF3 [65]. The resulting MOF crystals
exhibit identical cubic morphology and a very uniform distribution of particle
Luminescent Properties and Applications of Metal-Organic Frameworks 35

sizes, which can be varied from micrometer down to submicrometer scale by


controlling the reactant concentration. The uniform size and shape of MOFs are
attributed to the ability of the microwave technique to control the nucleation
process throughout the volume of solution, in which all of the crystals are nucleated
at once. Recently, the use of microwave-assisted synthesis to construct MIL-47 and
six new vanadium MOFs has also been reported by Centrone et al [66].

2.2 Nanoscale MOFs

Although MOFs have shown promise for a number of practical applications, the bulk
crystalline nature do not always fulfill all specific needs for these applications.
For example, in the fields of biology, drug delivery, and biomedical imaging, the
nanoscale dimensions are required to provide materials of sufficiently small sizes for
their internalization into cells [67]. Recently, scaling down these framework materials
has afforded an exciting new class of materials known as nano-MOFs (NMOFs)
[68–71], which exhibit the combination of the rich diversity of compositions,
structures, and properties of bulk MOFs with the obvious advantages of
nanomaterials. The limited solution-based behavior of bulk MOFs can be overcome
by the preparation of nano-MOFs, enabling the MOFs highly dispersed in aqueous
media or other solvents and coated for improving their biocompatibility. In addition,
the NMOFs may display the higher surface areas than their macroscopic counterparts
and the unique size-dependent optical, electrical, and magnetic properties.
Usually, NMOFs can be produced through two major strategies [67]. One is the
controlled precipitation of self-assembled MOFs once they are formed, which includes
the use of microwaves or sonochemistry methods. As mentioned above, the micro-
wave radiation and ultrasounds can accelerate the nucleation and increase the seed
number, avoiding the fast growth of MOF crystals. Therefore, microwave and
sonochemistry synthesis methods are the productive strategies generally used to
produce a wide variety of NMOFs. The other method is confining the assembly process
at specific nanoscale locations. For example, the supramolecular assembly of NMOFs
can be confined into droplets by using microemulsion method or controlled on the
surface of substrates by using template. Microemulsion is a suspension consisting of
small droplets of an immiscible phase (nonpolar or polar) dispersed in a continuous
phase, usually stabilized by a surfactant. Microemulsion system can be divided into
oil-in-water and water-in-oil, depending on the concentration of the different
components. By varying the concentration of the dispersed phase and the surfactant,
the size of the droplets can be tuned in the range 50–1,000 nm, approximately. These
nanodroplets in the microemulsion can be used as “nanoreactors” to control the
kinetics of particle nucleation and growth [72].
The utilization of water-in-oil, or reverse, microemulsion method for NMOFs syn-
thesis has been reported by Rieter et al. [73], who prepared a NMOF Gd(BDC)1.5(H2O)2
(BDC ¼ 1,4-benzenedicarboxylate) by reacting GdCl3 and bis(methylammonium)ben-
zene-1,4-dicarboxylate in the cationic cetyltrimethylammonium bromide (CTAB)/
36 Y. Cui et al.

isooctane/1-hexanol/water system. By varying the w value (defined as the water/surfac-


tant molar ratio) of the microemulsion systems, the morphologies and sizes of the
NMOFs can be tuned. Furthermore, the particle size of NMOF is also affected by the
reactant concentration and the reactant ratio. The crystalline nanoplates of [Gd(1,2,4-
BTC)(H2O)3]·H2O (BTC ¼ benzenetricarboxylate), with a diameter of ~100 nm and an
average thickness of 35 nm, were also prepared using the reverse microemulsion
method. This group also synthesized the NMOFs Mn(BDC)(H2O)2 and
Mn3(BTC)2(H2O)6 in reverse microemulsions [74]. The reaction of MnCl2 and
[NMeH3]2(BDC) in a CTAB/1-hexanol/n-heptane/water microemulsion with a
w value of 5 yielded the nanorods of Mn(BDC)(H2O)2, with a structure corresponding
to a known crystalline phase. NMOF of Mn3(BTC)2(H2O)6 can be produced in a CTAB/
1-hexanol/isooctane/water microemulsion (w ¼ 10) with a Na3(BTC)/MnCl2 molar
ratio of 2:3. The particles of Mn3(BTC)2(H2O)6 are crystalline spiral rods, but do not
correspond to any known phase of Mn-BTC-based MOFs. The similar synthesis of
luminescent NMOF Eu2(BDC)3(H2O)2·(H2O)2 by reacting EuCl3 and the
methylammoium salt of H2BDC in a CTAB/1-hexanol/isooctane/water microemulsion
was reported by Xu et al. [75].
Besides the room-temperature microemulsion synthesis, the surfactant-assisted
synthesis at elevated temperatures has been developed by the Lin group for the
synthesis of NMOFs [76]. The blocklike nanoparticles of Gd2(bhc)(H2O)6 (bhc ¼
benzenehexacarboxylate) were synthesized by heating a CTAB/1-hexanol/n-heptane/
water microemulsions containing [NMeH3]6(bhc) and GdCl3 at 120 C for 18 h.
Interestingly, the reaction of GdCl3 and mellitic acid in the same microemulsion
system at 60 C in a microwave reactor for 15 min afforded the high-aspect ratio
nanorods, which has a formula of Gd2(bhc)(H2O)8·2(H2O). Additionally, the
surfactant-assisted synthesis under microwave heating is also utilized to produce
the Mn-based NMOFs, which have been synthesized in reverse microemulsions at
room temperature, as mentioned above. Nanorods of Mn(BDC)(H2O)2 synthesized
under microwave heating have identical phase and similar morphology to those
obtained from the microemulsion, while the particles of Mn3(BTC)2(H2O)6 obtained
under microwave heating have a blocklike morphology, with lengths ranging from
~50 to 300 nm in the three dimensions [74].

2.3 Postsynthesis Modification

The decoration of the pores and modification of the structures afford the possibility
to modulate the physical and chemical properties of MOFs and tune the interactions
of guest species with host frameworks, enabling the MOFs useful for more
specialized and sophisticated applications. Modification and functionalization of
MOFs have been reported by using unsaturated metal centers or functional groups
on the organic bridging ligands [77–81]. Although the method of using modified
ligand directly in synthesis has resulted in MOFs with pendant groups such as Br,
NH2, and CH3 decorating the pores, it is limited because some kinds of
functional groups, such as OH, COOH, and N-donating groups, have a tendency
Luminescent Properties and Applications of Metal-Organic Frameworks 37

to coordinate to metal ions. In addition, the incorporation of functional groups on


the linking ligands may lead to insolubility, thermal instability, or steric bulk,
hindering the formation of desirable MOF structures. Postsynthesis modification
has recently emerged as a highly versatile tool to functionalize MOFs, which allows
introducing covalent attachment to the organic linker after the framework has been
synthesized. The realization of postsynthesis approach within MOFs has certainly
enriched the MOFs with predictable structures [82]. Incorporation of functional
organic and/or metal-organic sites within MOFs is not only very important to tune
the porosity and introduce their specific recognition of substrates, but also crucial to
systematically modulate their luminescent functionalities.
The MOF IRMOF-3, which is constructed from Zn4O SBUs and 2-amino-1,4-
benzenedicarboxylic acid (NH2–BDC) linkers, has been used to demonstrate the
covalent postsynthesis modification by Cohen et al. [83, 84]. This MOF can be
reacted with alkyl anhydrides to yield a series of amide-bearing MOFs with different
substituents, but maintain the same crystallinity and thermal stability of IRMOF-3.
This group also incorporated a series of medium to long alkyl groups within IRMOF-
3 through postsynthesis method to yield the hydrophobic MOFs [85]. In the same
study, the MOF MIL-53(Al)–NH2 was also modified with three different alkyl
anhydrides to generate MIL-53(Al)–AM1, MIL-53(Al)–AM4, and MIL-53
(Al)–AM6. The MOF MIL-53(Al)–NH2 and MIL-53(Al)–AM1 are hydrophilic
with contact angles of 0o, while MIL-53(Al)–AM4 and MIL-53(Al)–AM6 show
superhydrophobic properties with contact angles greater than 150 . This was pro-
posed that the combination of the submicrometer crystallite size and the hydrophobic
functional groups result in the superhydrophobicity. The similar modification was
also executed using the MOFs IRMOF-3, DMOF-1–NH2, and UMCM-1–NH2 to
enhance hydrogen sorption properties of MOFs [86]. Taylor–Pashow et al. have also
utilized the covalent postsynthesis method to attach drugs and imaging contrast
agents to a Fe-based MOF, enabling it suitable for imaging and therapeutic
applications [87].
The utilization of “click” chemistry for postsynthesis modification of MOFs has
been reported by Goto et al. [88], who reacted a MOF bearing the azide group in the
organic linkers with some alkynes containing various functional groups through a
copper(I)-catalyzed azide–alkyne cycloaddition reaction. Savonnet et al. developed
a one-pot, two-step modification method starting from already available amino-
derived MOFs, DMOF–NH2 [Zn(bdc–NH2)(DABCO)] and MIL-68(In)–NH2 [In
(OH)(bdc–NH2)] [89]. In this method, the amino groups of DMOF–NH2 were
transformed into an azide DMOF–N3 and then reacted with an excess of
phenylacetylene and Cu(CH3CN)4PF6 via “click” reaction to generate the triazole-
modified DMOF. The modification was also executed for the MIL-68(In)–NH2, and
the same results was obtained. A method to selectively modify both the surface and
interior of a MOF has been developed by Gadzikwa et al. [90], who prepared a
non-catenated, pillared paddle-wheel framework TO-MOF using 1,2,4,5-tetrakis
(4-carboxyphenyl)benzene (TCPB), 3-[(trimethylsilyl)ethynyl]-4-[2-(4-pyridinyl)
ethenyl]pyridine, and Zn2+ in acidic DMF at 80  C for 1day. The trimethylsilyl
(TMS) groups within To-MOF can be easily removed via deprotection, and the
38 Y. Cui et al.

exposed terminal alkyne groups can react with benzyl azide to form a triazole via the
“click” reaction. In order to perform a selective modification, To-MOF was firstly
exchanged with CHCl3 and then exposed in an aqueous solution of KF to yield the
surface-deprotected MOF, because the KF cannot enter the pores which are filled
with CHCl3. Thus, only the deprotected surface of MOF can be reacted with the dye
ethidium bromide monoazide to obtain the surface-modified MOF. Subsequently,
the interior of the MOF can be deprotected using triethylammonium fluoride (TEAF)
and further reacted with benzyl azide to achieve the differentially
functionalized MOF.

3 Luminescent Properties of MOFs

So far, hundreds of luminescent MOFs have been reported, and the luminescence can
exist in several forms. Luminescence can arise from direct organic ligands excitation
(particularly from the highly conjugated ligands), metal-centered emission (widely
observed in lanthanide MOFs through the so-called antenna effect), and charge
transfer such as ligand-to-metal charge transfer (LMCT) and metal-to-ligand charge
transfer (MLCT). Furthermore, the guest molecules can also result in luminescence
onto MOFs.

3.1 Ligand-Based Luminescence

Generally, when an organic molecule absorbs a photon of appropriate energy, a


chain of photophysical events occur, including internal conversion or vibrational
relaxation, fluorescence, intersystem crossing, and phosphorescence. These pro-
cesses can be divided into radiative and nonradiative processes. Nonradiative pro-
cesses include internal conversion, vibrational relaxation, and intersystem crossing.
Internal conversion is defined as the nonradiative transition between different elec-
tronic states of the same multiplicity, whereas intersystem crossing is defined as that
between states of different multiplicity. Vibrational relaxation is a process in which
the content of the vibrational energy in a given electronic state changes. This process
includes vibrational-energy deactivation and vibrational-energy redistribution. The
former process accompanies energy loss of the excited molecules, whereas the latter
is normally an isoenergetic process in which excess energy in given vibrational
modes redistributes to other modes. Radiative processes of organic molecules
include fluorescence and phosphorescence emission. Fluorescence is defined as a
radiative transition between electronic states of the same multiplicity, such as the
radiative transition from its first singlet state S1 to the ground singlet state S0, which
is spin allowed. Phosphorescence of organic molecules corresponds to the radiative
transition from triplet state T1 to S0, which is spin forbidden and has lifetime on the
order of several microseconds to seconds.
The luminescence properties can be characterized by the following parameters:
(1) Luminescence spectra, defined as fluorescence intensity as a function of a
Luminescent Properties and Applications of Metal-Organic Frameworks 39

wavelength; (2) Quantum yield, the ratio of the total number of emitted photons
released in the process of fluorescence to the total number of molecules promoted to
the excited state. The emission quantum yield of all emissive organic molecules is <1.
This observation implies that there are nonradiative processes other than radiative
processes. Further, it was revealed that nonradiative processes are essentially intra-
molecular in nature even under the collision-free conditions; and (3) Lifetime, which
refers to the average time the molecule stays in its excited state before emitting a
photon and is determined as being inversely proportional to the sum rate constants
of a radiative process and the nonradiative processes collectively known as
quenching [91].
The ligands in MOFs are commonly π-conjugated organic molecules due to their
rigidity, and the majority of them are based on rigid backbones functionalized with
multi-carboxylate groups or heterocyclic groups for metal–ligand coordination.
Usually, the fluorescence emission of organic ligands is similar to their emission
behavior in solution, corresponding to the transition from the lowest excited singlet
state to the singlet ground state, and the transitions are either π ! π* or n ! π* in
nature. However, the fluorescence properties such as maximum emission wave-
length and lifetime of organic linkers in solid MOFs are often different from those
of the free molecules. This is because the organic linkers are stabilized within
MOFs, which reduces the nonradiative decay rate and leads to increased fluores-
cence intensity, lifetimes, and quantum efficiencies. In the solid state, molecular
interactions make the molecules close together, which enable charge transfer
among the organic ligands/linkers, resulting in shift of spectra, broadening of the
emission, and loss of fine structure. In addition, the size and nature of metal ions,
the orientation and arrangement of linkers, and the coordination environment
within MOF can affect the fluorescence properties of the organic linkers because
these factors will induce their different intramolecular or intermolecular
interactions among organic linkers. Therefore, controlling these interactions is
crucial to tune the luminescence properties of MOFs for a particular application.
The following examples show how intramolecular/intermolecular interactions of
organic linkers have played the important roles on the luminescence properties of
the resulting MOFs.
The MOF Zn3(μ5-pta)2(μ2-H2O)2 (pta ¼ 2,4,6-pyridinetricarboxylate) exhibits
strong luminescence at 467 nm [92], while the free ligand molecule displays a
weak luminescence at 415 nm when excited at 338 nm at room temperature.
The fluorescent enhancement and red shift of organic linkers within this MOF
is attributed to the formation of the framework structure which has enabled the
rigidity of the aromatic backbones and maximized the intramolecular/intermolecular
interactions among the organic linkers for their energy transfer and decreased the
intraligand HOMO–LUMO energy gap. Coordination-perturbed ligand-centered
luminescence was reported in MOFs Zn3(BTC)2(DMF)3(H2O)·(DMF)(H2O) and
Cd4(BTC)3(DMF)2(H2O)2·6H2O (BTC ¼ 1,3,5-benzenetricarboxylate) [93]. The
strongest emission at 370 nm of the free BTC when excited at 334 nm, attributed
to the π* ! n transitions, was shifted to 410 and 405 nm in these two MOFs,
respectively, when excited at 341 and 319 nm.
40 Y. Cui et al.

Usually, most organic molecules are highly emissive in solution but become
weak or even totally annihilated when the molecules are aggregated due to energy
transfer and the formation of excimers and exciplexes. However, some molecules
can exhibit the opposite effect: they show no emission when dissolved in good
solvents but become highly emissive upon aggregation in poor solvents or in the
solid state. This process is termed as “aggregation-induced emission (AIE),” which
makes it possible to actively utilize the aggregation process, instead of passively
working against it. Recently, Dinca et al. describe the new coordination-induced
emission effect (MCIE) through coordinative immobilization of tetrakis
(4-carboxyphenyl)ethylene (TCPE) within rigid porous MOFs [94]. Although
with the absence of close-packed tetraphenylethylene cores, which are required
for fluorescence “turn-on” in AIE, the MOFs Zn2(TCPE)(H2O)2·4DEF and
Cd2(TCPE) (DEF)(C2H5OH)2·DEF exhibit strong emission at 480 and 455 nm,
respectively, similar to the value observed for solid H4TCPE, whose emission
maximum is 480 nm (Fig. 1). Furthermore, both the MOFs exhibit biexponential
fluorescence decays with values similar to those observed for solid H4TCPE. This
result demonstrates that anchoring AIE-type molecules to metal ions within a MOF
is also an effective method to turn-on fluorescence by restricting the rotation of the
phenyl rings.

3.2 Lanthanide Luminescence

Lanthanide ions (Ln3+) are characterized by a gradual filling of the 4f orbitals, from
4f0 (for La3+) to 4f14 (for Lu3+). The valence 4f electrons of trivalent lanthanide ions,
Ln(III), are well shielded from the environment by the outer core 5s and 5p electrons
and are minimally involved in bonding [95]. Because of this shielding, the atomic
properties of these ions are typically retained after complexation. The absorption and
emission spectra of Ln(III) ions consist of sharp, narrow bands corresponding to the
f–f transitions of the metal ion. Figure 2 shows the most significant energy levels for
several lanthanide ions. The f–f transitions are Laporte forbidden, resulting in low
molar absorption coefficients in the near-UV and visible ranges (<10 M1 cm1)
and long luminescence lifetimes in the millisecond range. The number of observed
bands depends on the particular lanthanide ion and its arrangement of electronic
states. Except for lanthanum (La3+) and lutetium (Lu3+), each lanthanide ion has its
specific emission bands: for example, the lanthanide ion europium (Eu3+) emits red
light; terbium (Tb3+) green light; and neodymium (Nd3+), ytterbium (Yb3+), and
erbium (Er3+) near-infrared light.
The lanthanide ions suffer from weak light absorption due to the forbidden f–f
transitions, making the direct excitation of the metals very inefficient unless high-
power laser excitation is utilized. This problem can be overcome by coupling
species that can participate in energy transfer processes, known as “luminescence
sensitization” or “antenna effect” [96–98]. The mechanism of antenna sensitization
within MOFs is comprised of three steps: light is absorbed by the organic ligands
around the lanthanide ions; energy is transferred to the lanthanide ions from organic
Luminescent Properties and Applications of Metal-Organic Frameworks 41

a b
455 480

normalized emission / absorption


fluorescence by non-emissive
molecular aggregation TPE rotor
(AIE)
fluorescence by coordination in rigid matrix
(MCIE) 300 400 500 600 700
λ / nm

Fig. 1 (a) Turn-on fluorescence in a TPE rotor by aggregation (AIE) and by coordination in a
rigid MOF matrix (MCIE) and (b) Diffuse reflectance and emission spectra of H4TCPE (blue),
Zn2(TCPE)(H2O)2·4DEF (red), and Cd2(TCPE) (DEF)(C2H5OH)2·DEF (green). Reprinted with
the permission from [94], Copyright 2011, American Chemical Society

35000

6
P7/2
30000 1
D2

5
25000 G4
5
D2 1
G4
5
Energy (cm-1)

D4
5
20000 3
P0 D1 4
F9/2 5
S2
1
D2
4 5
G5/2 5 F4 4
F9/2
D0
15000 6
F3/2 5
I4 4 3
3
F3.2
I9/2 H4
5
1 5
6
F11/2 I5 4
2
F5/2
G4 4
F1 I11/2
10000 F3/2 5
I6 3
H5
3 5
F4 IS 4
4
I15/2 7 I13/2 3
F4
7 F4 5
5I F6 I7
4
5000 I13/2 7 6
H13/2
2 3
F7/2 H5 4
I11/2

0
Ground = 2
F5/2
3
H4
4
I9/2
5
I4
6
H5/2 7
F0 5
S
7
F4 6
H15/2 5
IS 4
I15/2
3
H6 2
F7/2
State
La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu
n=0 1 2 3 4 5 6 7 8 9 10 11 12 13 14

f n electron configuration (n = 0-14)

Fig. 2 A summary of electronic excited-state energy levels for Ln3+ ions. Reprinted with the
permission from [95], Copyright 2009, American Chemical Society

ligands; and then luminescence is generated from the lanthanide ions. One of the
main energy migration paths is ligand-centered absorptions followed by inter-
system crossing S1 ! T1, T1 ! Ln3+ transfer, and metal-centered emission.
Another possible path is the direct transfer of energy from the excited singlet
42 Y. Cui et al.

state S1 to the energy levels of the lanthanide ion, which is known for Eu3+ and
Tb3+. Because of the energy transfer of the organic ligands to the lanthanide ion, the
fluorescence and phosphorescence of the ligands is quenched. If such energy
transfer is not very efficient, both remaining ligand fluorescence and the
lanthanide-centered luminescence could be observed. Furthermore, ligand-to-
metal charge transfer (LMCT), metal-to-ligand charge transfer (MLCT), and
4f–5d transitions may also funnel energy onto the lanthanide ions. For trivalent
lanthanide ions such as Sm3+, Eu3+, and Yb3+ that can easily be reduced to the
divalent ions, the excitation energy can be transferred from an LMCT state to the 4f
levels of the lanthanide ions when the LMCT state lies at high-enough energy.
Detailed investigation on the numerous energy transfers processes is very important
in order to well tune the luminescent properties of lanthanide MOFs.
In MOFs, efficient lanthanide-centered luminescence is usually fulfilled by the
use of antenna linkers, in which, the following conditions have to be considered:
(1) The intersystem crossing yield of the antenna linkers should be high, although
the lanthanide ion can enhance the intersystem crossing yield via an external heavy
atom effect. (2) The linkers should have matching triplet state energy levels for the
resonance level of the lanthanide ion. If the energy difference between donor and
acceptor is too small, a thermally activated energy back transfer can occur, whereas
large energy differences may lead to slower energy transfer rates. The energy of the
triplet state must be elaborately tuned to maximize the transfer and minimize the
back transfer. (3) The antenna linkers should have a high absorption coefficient, in
order to obtain high luminescence intensities. (4) The antenna linkers should be in
close proximity to the lanthanide ion, because the energy transfer process is
strongly distance dependent.
Thus, it is of great significance to select or design suitable ligand with the
appropriate energy level. To illustrate the sensitizations effect of ligands, we show
here two representative examples of such luminescent lanthanide MOFs. The first
example is the isostructural lanthanide (Sm3+, Eu3+, Gd3+, Tb3+, Dy3+) MOFs
containing 4,40 -disulfo-2,20 -bipyridine-N,N0 -dioxide as linker. The triplet state
energy levels of ligand was determined to be 24,038 cm1; thus the ligand is capable
of sensitizing the trivalent ions Sm3+, Eu3+, Tb3+, and Dy3+ in which strong
emissions observed have been observed in Eu3+ and Tb3+ MOFs. In the second
example, the thiophenyl-derivatized carboxylic acid has been selected as possible
sensitizing chromophore and multicarboxylate linker to construct the lanthanide
MOFs [99]. Because of the big radius of the S atom, its lone pair of electrons can be
more easily delocalized within the heterocycle, and the ligand exhibits good charge-
transfer ability. The results show that Eu3+ and Tb3+ MOFs exhibit typical intense
luminescence in the visible region, and the Yb3+ MOFs emit moderate characteristic
near-infrared luminescence in solid state even with coordinated solvents.
In hetero-lanthanide MOFs, a similar energy transfer from ligands to lanthanide
occurs through which multiple lanthanide elements may also be excited, but instead
of stimulating emission from each, energy can be transferred from one lanthanide to
another, resulting in a preferential enhancement of a single lanthanide-centered
luminescence. For instance, the MOF [(Eu, Tb)(C6H8O4)3(H2O)2]·(C10H8N2)
Luminescent Properties and Applications of Metal-Organic Frameworks 43

shows the three major peaks at 545, 595, and 615 nm resulting from both
Tb3+ and Eu3+ emission [100]. The peaks corresponding to Eu emission (595 and
615 nm) are more intense than in the isostructural Eu-only analog, and Tb emission
at 545 nm is nearly completely quenched, suggesting that the Tb3+ center is
also acting to sensitize the Eu3+ then effectively enhancing Eu3+ emission. The
similar Tb3+ to Eu3+ energy transfer has also been observed in the MOFs
(Eu0.2Tb0.8)2(2,5-pdc)2(1,4-pda)(H2O)2 and (Eu0.1Tb0.9)2(2,5-pdc)2(1,4-pda)
(H2O)2 (H2pdc ¼ 2,5-pyridinedicarboxylic acid, H2pda ¼ 1,4-phenylenediacetic
acid) [101]. In these MOFs, the 14 K emission spectra excited at the Tb3+ 5D3 level
(377 nm) show the typical Eu3+ and Tb3+ lines, clearly supporting the
abovementioned Tb3+ to Eu3+ energy transfer. Excitations at the band from
255 to 310 nm produce similar spectra to those obtained under direct Tb3+ excita-
tion, suggesting that ligand to Tb3+, ligand to Eu3+, and Tb3+ to Eu3+ energy
transfer occurs. On the contrary, the spectra excited at the Eu3+ 5D2 level
(464 nm) only show the typical Eu3+ lines.
Up-conversion is the generation of higher-energy photons from lower-energy
radiation on the basis of sequential adsorption and energy-transfer steps.
Up-conversion luminescent materials have been extensively investigated for their
potential applications in laser, display, bioassay, and bioimaging. In particular,
NIR-to-visible up-conversion luminescent materials which emit visible light upon
NIR excitation have attracted much attention in biomedicine due to the absence of
photodamage to live organisms, low autofluorescence background, high signal-to-
noise ratio, and high light-penetration depth in biological tissues [102, 103]. Up to
now, many kinds of lanthanide ion-doped inorganic materials have been widely
investigated for up-conversion luminescence. However, only a few examples of
up-conversion luminescent lanthanide MOFs have been reported. Yang
et al. prepared a MOF Nd2(1,4-NDC)3(DMF)4·H2O (1,4-NDC ¼ 1,4-naphthalene-
dicarboxylate) [104], which exhibits a weak UV up-conversion emission at about
391.6 nm and a much stronger blue emission at about 449.5 nm upon pulse laser
excitation at 580 nm. Weng et al. reported a Y:Er-Yb codoped MOF [(Y:Er-
Yb)3(1,4-BDC)3.5(OH)2(H2O)2]·H2O [105], which exhibits four emission bands
of Er3+ under 980-nm laser excitation. The red emission attributed to the
4
F9/2 ! 4I15/2 transition is centered at 654 nm, and the green bands ascribed to
the transitions of 4S3/2/2H11/2 ! 4I15/2 are at 545 and 524 nm, respectively. A rarely
observed indigo emission at 455 nm can also be observed, corresponding to the
transition 4F5/2 ! 4I15/2. The 545- and 524-nm emissions are described as a
two-photon up-conversion excitation mechanism, and the up-conversion process
is also discussed. First, the Yb3+ ions are easily excited to their 2F5/2 level by
absorbing a 980-nm photon, and then the energy is transferred to the Er3+ ions
which can be excited to the 4I11/2 level. A second photon transfer from the excited
Yb3+ ions pumps the Er3+ ions from 4I11/2 to the 4F7/2 level, which can decay
nonradiatively to the 4S3/2 and 2H11/2 levels. Thus, the green emissions observed
result from the 4S3/2 ! 4I15/2 and the 2H11/2 ! 4I15/2 transitions. Furthermore, the
455-nm emission is explained through the three-photon up-conversion excitation
mechanisms.
44 Y. Cui et al.

The luminescence of Ln3+ ion from the f–f transitions can be classified as two
types of transitions: the parity-allowed magnetic dipole transitions and the parity-
forbidden electric dipole transitions. When the Ln3+ ion is inserted into a chemical
environment, noncentrosymmetric interactions allow the mixing of electronic states
of opposite parity into the 4f wave functions, and electric dipole transitions become
partly allowed. The intensity of some of these transitions is particularly sensitive to
the nature of the metal ion environment, and these transitions are called hyper-
sensitive transition; a typical example is the 5D0 ! 7F2 transition of Eu3+
[106]. Thus, the luminescence of lanthanide ions can provide valuable information
about the local environment and make them very suitable for acting as a structural
probe deciphering the symmetry of the chemical environment and the coordination
sphere.
Furthermore, the lanthanide-centered emission in MOFs is sensitive to the lowest
triplet level of the ligands, suggesting the possibility to modulate the luminescence
intensity for a given lanthanide ion by varying the interaction between the ligands
and the analyte. These interactions may facilitate or disrupt the energy transfer
process, including a modification of the energy transfer ability of the bound ligands
or energy transfer from the analyte onto the lanthanide ion, even quenching instead
of sensitization. Therefore, the luminescence of lanthanide ions is also potentially
useful for a variety of analytical probes. Nowadays, chemical sensors based on
lanthanide MOFs for cations, anions, small molecules, gases, and explosives have
been designed along these lines.
As remarked above, the luminescence of Eu3+ ions is most appropriate as
structural probe for the determination of the number of metal ion sites in a
compound, their symmetry, their hydration numbers, and their coordination sphere.
Especially, the ratio of the intensities of the 5D0 ! 7F1 and 5D0 ! 7F1 transitions
is very sensitive to the symmetry of the Eu3+ centers, because the 5D0 ! 7F1
emission is due to the magnetic dipole and independent of the environment, while
the 5D0 ! 7F2 emission is due to the electric dipole and is sensitive to the crystal
field symmetry. Four examples illustrating the luminescence of Eu3+ ions can be
utilized as a structural probe in MOFs. In the first example, a Eu-doped MOF
[La2(H2O)4][(C5H3N(COO)2)2(C6H4(COO)2)] [107], the intensity ratio of the
5
D0 ! 7F2 and 5D0 ! 7F1 transitions is 5.2, and the intensity ratio of dehydrated
samples is determined to be 4.2. The fully hydrated MOF contains two water
molecules positioned adjacent to each other in one face of the trigonal prism, as
part of the nine coordination of the La3+ ion, and can exert some strain. Upon
dehydration, the bound water molecules are removed, which would give rise to a
seven-coordinated La3+ ion that is probably much less strained than the nine-
coordinated La3+ site. Thus, the more symmetric environment of lanthanide ions
in the dehydrated MOF results in the decrease of the intensity ratio value.
The second example MOF Eu3(2,6-pydc)3(2,6-Hpydc)(SO4)(H2O)3·3(H2O)3
(2,6-pydc ¼ pyridine-2,6-dicarboxylate) [108] crystallizes in the triclinic space
group and possesses a 2D metal-organic framework based on hexanuclear Eu6
SBUs. The asymmetric unit comprises three independent Eu3+ ions, three pydc2
dianions, one Hpydc anion, one sulfate, three coordinated, and three guest water
Luminescent Properties and Applications of Metal-Organic Frameworks 45

molecules. The intensity ration of the 5D0 ! 7F2 and 5D0 ! 7F1 transitions is
about 5.2; in addition, the symmetry-forbidden emission 5D0 ! 7F0 is observed
at 579 nm. These phenomena indicate that Eu3+ ions occupy sites with low
symmetry and have no inversion center.
In the third example, Chandler et al. synthesized a MOF [Na6(H2O)6][Eu(L)4]
(H2O)nCl (L ¼ 4,40 -disulfo-2,20 -bipyridine-N,N0 -dioxide) [109], in which, the
metal center exists in a mildly distorted square antiprismatic geometry resulting
in a D4 point symmetry based on the crystallographic data. Generally, when Eu3+
ions exist in a D4 site symmetry, two components of the 5D0 ! 7F1 and one
component of the 5D0 ! 7F2 transitions are expected to be observed, and the
5
D0 ! 7F0 transition remains strictly forbidden. While the luminescence spectrum
displays a weak band ascribed to 5D0 ! 7F0 transition as well as two peaks of the
5
D0 ! 7F1 and two well-resolved sharp peaks of the 5D0 ! 7F2 transition,
indicating a small distortion from an ideal D4 symmetry. When the MOF is fully
dehydrated via thermal treatment, the emission spectrum remains nearly identical to
the original, except the weak band of 5D0 ! 7F0 transition disappears, suggesting
that the water can perturb the ligand system to a minor extent inducing the
noncentrosymmetric ligand field.
Finally, the emission spectra and 5D0 lifetime of Eu3+ are utilized to estimate the
strength of the Ln–ligand bond and the hydration numbers. Shi et al. describe a Eu
MOF linked with etidronic acid, named as Na4[Eu2(H2O)2(hedp)2]·H2O [110],
whose structure comprises eight-membered ring channels filled with Na+ and
both free and Eu-coordinated water molecules. When excited at 250 and 395 nm,
the luminescence spectra display the typical Eu3+-centered emission ascribed to the
5
D0 ! 7F0-4 transitions, whose energy and number of Stark components is inde-
pendent of the excitation wavelength, and only one line for the 5D0 ! 7F0 transi-
tion is observed. This suggests that the Eu3+ ions occupy the same average local
environment in as-synthesized and dehydrated MOF. Dehydration induces the
decrease and red shift for the 5D0 ! 7F0 transition, indicating that the Eu–O
bonds are slightly less covalent in the dehydrated MOF, in accord with the X-ray
diffraction evidence. Additionally, the presence of water molecules in the first
coordination sphere is estimated from the emission spectra and 5D0 lifetime values.
As mentioned above, lanthanide luminescence is very sensitive to the quenching by
high-energy vibrations such as O–H. This quenching can be turned into an advan-
tage for assessing the number of water molecules q interacting in the inner coordi-
nation sphere from lifetimes measured in water and deuterated water. Assuming
that the O–H quenching is the main nonradiative process operating and that all the
other deactivation paths are the same in water and in deuterated water, the hydration
number can be determined by measuring the lifetime in the deuterated solvent
[106]. Thus, the number of water molecules coordinated to Eu3+ is determined
using an empirical formula, and the results for the as-synthesized and rehydrated
samples are very similar, indicating one water molecule in the coordination sphere
of Eu3+, in accord with the crystal structure.
46 Y. Cui et al.

3.3 Charge-Transfer Luminescence

Charge-transfer luminescence is an allowed transition from the charge-transfer


excited state to the ground state. Ligand-to-metal charge transfer (LMCT) and
metal-to-ligand charge transfer (MLCT) states are the excited states commonly
found in MOFs. LMCT states involve the electronic transitions from a ligand-
localized orbital to a metal-centered orbital, while MLCT states correspond to the
electronic transitions from a metal-centered orbital to a ligand-localized orbital.
The charge-transfer luminescence is commonly observed in d10 metal-based MOFs.
For example, in the MOFs Cu3(C7H2NO5)2·3H2O (C7H2NO5 ¼ 4-hydroxy-
pyridine-2,6-dicarboxylate) and CuAg2(C7H3NO5)2 [111], the former exhibits
blue fluorescence with the maximum emission at 398 and 478 nm upon excitation
at 333 nm, while the latter and the free ligand 2,6-dicarboxy-4-hydroxypyridine
both display green fluorescence with the maximum emission at 515 and 526 nm
upon excitation at 358 and 365 nm, respectively. Compared with the ligand, the
MOF Cu3(C7H2NO5)2·3H2O displays two large blue shift of 59 and 48 nm in
luminescent spectra, respectively, while CuAg2(C7H3NO5)2 displays a little blue
shift of 11 nm, suggesting that the emission of the former may be originated from
MLCT. The origin of the emission bands are further confirmed by the calculated
energy band structures. This indicates that the luminescence of MOF
Cu3(C7H2NO5)2·3H2O may be ascribed to MLCT where the electrons are trans-
ferred from the Cu-3d to O-2p and N-2p states, while the emission band of
CuAg2(C7H3NO5)2 may be originated from π–π* transition of the ligand.
Luminescence originated from LMCT excited states have been reported in some
MOFs. For example, the MOF [Zn(2,3-pydc)(bpp)]·2.5H2O and [Cd(2,3-pydc)
(bpp)(H2O)]·3H2O (2,3-pydcH2 ¼ pyridine-2,3-dicarboxylic acid) display intense
fluorescent emissions at 436 and 438 nm upon excitation at 372 and 370 nm [112],
respectively, while the ligand 2,3-pydcH2 displays very weak photoluminescent
property upon excitation at 370 nm. Thus the origin of the emission for these MOFs
can be attributed to LMCT luminescence. In a homochiral Cd MOF [Cd(dtba)
(bpp)]3n (H2dtba ¼ 2,20 -dithiobisbenzoic acid, bpp ¼ 1,3-bis(4-pyridyl)propane)
[113], the ligands H2dtba and bpp both act as multidentate bridging ligands to
improve the chain helicity of the product and have the ability to produce unique
interwoven extended structural motifs. Interestingly, the luminescence of the MOF
exhibits clear difference when excited at room and low temperatures. The MOF
displays a blue emission band at 434 nm with a shoulder peak at 482 nm when
excited at 355 nm at room temperature, which is attributed to the metal-perturbed
intraligand emissions of dtba or bpp. When frozen to 10 K, the emission of the MOF
shifts to longer wavelength at about 507 nm and with only one peak, which is
attributable to LMCT.
Sometimes, LMCT or MLCT luminescence may compete with ligands-based
luminescence, resulting in two emission bands. For example, excitation of the zinc
MOF Zn2(ATA)3(ATA)2/2 at 320 nm at room temperature produces a blue lumines-
cence with a weak emission peak centered at 485 nm that may be assigned to the
Luminescent Properties and Applications of Metal-Organic Frameworks 47

LMCT or MLCT and a significantly strong emission with the maximum peak at
392 nm originated from ligands-based emission [114].

3.4 Guest-Induced Luminescence

Serving as host matrices, MOFs can offer a unique platform to incorporate and
confine the luminescent species due to their highly regular channel structures and
controllable pore sizes. In fact, through the introduction of a variety of metal ions,
quantum dots, nanoparticles, and fluorescent dyes into their pore spaces, many types
of guest-induced luminescence behaviors have been observed in the MOF
composites. An et al. prepared a series of lanthanide ion-doped MOFs Ln3+@bio-
MOF-1 (Ln3+ ¼ Tb3+, Sm3+, Eu3+, or Yb3+) from the as-synthesized bio-MOF-1 via
cation exchange process [115]. When excited at 365 nm, the doped MOFs emitted
their distinctive colors (Eu3+, red; Tb3+, green; Sm3+, orange pink), which are readily
observed with the naked eye. Tb3+@bio-MOF-1, Sm3+@bio-MOF-1, Eu3+@bio-
MOF-1, and Yb3+@bio-MOF-1 exhibit the emission at 545, 640, 614, and 970 nm,
respectively. It needs to be mentioned that there exists another main emission at
340 nm in all these lanthanide ion-doped MOFs, suggesting that energy migrates
through the same electronic levels located in the MOF chromophoric structure for
doped lanthanide ions. Notably, the characteristic luminescence of lanthanide ions
can even be detected in aqueous environments, despite the strong quenching effect of
water molecules, indicating that the bio-MOF-1 scaffold can not only effectively
sensitize but also sufficiently protect the lanthanide ions. Furthermore, the quantum
yields of lanthanide emission are all reasonably high in aqueous environments,
demonstrating that the lanthanide ions are well protected within the pores, and the
energy transfers from MOFs to lanthanide ions are efficient. Luo et al. prepared the
Eu3+- and Tb3+-doped MOFs which exhibit tunable luminescence properties and
sensing functions for metal ions [116].
Except for the lanthanide ions, complex, fluorescent dyes and quantum dots have
also been encapsulated into MOFs to generate or tune the luminescence. A
mesoporous MOF, [Tb16(TATB)16(DMA)24]·(DMA)91(H2O)108 (H3TATB ¼ 1,3,5-
tribenzoic acid), containing cages of 3.9 and 4.7 nm in diameter [117], emits strong
green light originated from Tb3+ at 488, 541, 584, and 620 nm. Upon ferrocene is
introduced into the framework by a sublimation procedure at 100 C, no strong green
emission is observed; instead, a weak and broad emission from the included ferrocene
molecules appears. This may be attributed to the efficient energy transfer from the host
framework to the ferrocene molecules. Elemental analysis of the material reveals
uptake of roughly 65 ferrocene molecules per formula unit or 4,420 per pore. Removal
of the ferrocene under vacuum and high temperature leads to recovery of the green
emission from the MOF; moreover, the spectrum exactly matched that of the
evacuated host crystals, and the ferrocene emission disappeared.
The incorporation of highly luminescent core–shell quantum dots (QDs) within
MOF-5 is achieved [118]. Through appropriate surface functionalization, the yellow-
and red-emitting QDs are solubilized within MOF-5 growth media. This permits the
48 Y. Cui et al.

Fig. 3 (a) Digital photograph of vials containing yellow (left) and red (right) QD@MOF-5
samples illuminated with a black light. The samples are kept in dichloromethane. (b) Linear
emission spectra of yellow and red QDs in solution (continuous lines) and of the correspondent
QD@MOF-5 samples (dots). Reprinted with the permission from [118], Copyright 2012,
John Wiley & Sons Ltd.

incorporation of the QDs within the evolving framework during the reaction. The
resultant MOF composites exhibit yellow emission centered at 578 nm and the red
emission at 637 nm, respectively (Fig. 3). The unchanged emission peak profile of the
QDs embedded within the MOF framework suggests that the harsh chemical environ-
ment during MOF formation did not degrade the QD core–shell structure.
In addition, the classical fluorescent dye rhodamine 6G (Rh6G) has been
encapsulated into a large porous MOF, Cd3(bpdc)3(DMF)]·5DMF·18H2O
(H2bpdc ¼ 4,40 -biphenyldicarboxylic acid), reported by Fang et al. Rh6G is a
well-known xanthene derivative used as lasers dye, and it exhibits strong absorption
in the visible region and a very high fluorescence quantum yield. The MOF Cd
3(bpdc)3(DMF)]·5DMF·18H2O possesses a 1D hexagonal nanotube-like channels
of 24.5  27.9 Å, allowing the entry of the Rh6G dye molecules. The Rh6G-filled
MOF exhibits a strongest emission peak at 563 nm originating from the dye Rh6G,
which is similar to the previous results obtained on Rh6G-doped materials [119].
Up to date, a variety of luminescent MOFs have been prepared using lanthanide
and transition-metal ions. According to the metal ion, three main types of luminescent
MOFs can be clearly differentiated: lanthanide-based MOFs, transition-metal-based
MOFs, and heterometal-organic frameworks. In addition, a few examples of bismuth-,
lead-, and indium-based MOFs have been also reported. The survey of luminescent
MOFs and their luminescence wavelength is summarized in Table 1.

4 Applications of Luminescent MOF

As mentioned before, the metal centers, organic moieties, metal-organic charge


transfer, and guest molecules within porous MOFs all can potentially generate
luminescence. Such uniqueness can allow us to generate luminescent MOF
materials with systematically varied emission wavelengths which are crucial for
Table 1 Summary of luminescent MOFs
MOF Emission wavelength (nm)Bermudez References
Lanthanide-based MOFs
La [La(pmtz)(TzC)(H2O)3](H2O) 530 [120]
La2(PDA)3(H2O)5 408 [121]
Ce [Ce2(pydc)2(μ4-SO4)·5H2O]·2H2O 357, 480 [122]
Nd Nd2(1,4-NDC)3(DMF)4·H2O 391.6, 449.5 [104]
[Nd4(ox)4(NO3)2(OH)2(H2O)2]·5H2O ~1,060 [123]
Nd(tta)3(μ-bpm)·MeOH ~440, ~910, ~1,070 [124]
[Nd(trans-DAM)2(H2O)2]ClO4·3H2O 912, 1,059, 1,333 [27]
[Nd(trans-DAM)2(H2O)2]NO3·3H2O 912, 1,059, 1,333 [27]
[Nd(trans-DAM)(cis-DAM)(H2O)2]Cl·5H2O 912, 1,059, 1,333 [27]
[Nd(trans-DAM)2(H2O)2]Cl·5H2O 912, 1,059, 1,333 [27]
[Nd(H2TETA)]NO3·2H2O ~910, ~1,060, ~1,340 [125]
Sm [Sm(pta)(H2O)5]·4H2O 558, 592, 639 [126]
[Sm4(pydc)2(μ4-C2O4)4·8H2O]·6H2O 468, 566, 628 [122]
Sm2(ATPA)3(DMF)2(H2O)2 450, 561, 597, 644 [127]
Eu Eu(1,4-BDC)3(H2O)4 592, 617 [128]
Eu2(1,4-BDC)3(DMF)2(H2O)2 596, 616 [129]
[Eu(pta)(H2O)5]·4H2O 589, 613, 695 [126]
[Eu2(adipic acid)3(H2O)2]·4,40 -dipyridyl 595, 615 [130]
Eu3(2,6-pydc)3(2,6-Hpydc)(SO4)(H2O)3·3(H2O)3 579, 592, 614, 651, 696 [108]
Luminescent Properties and Applications of Metal-Organic Frameworks

Eu(PDC)1.5(DMF)(DMF)0.5(H2O)0.5 590, 616, 698 [131]


Eu2(FMA)2(OX)(H2O)4·4H2O 591, 616, 650, 692 [132]
Eu(BTC)(H2O)·1.5H2O 590, 616, 698 [133]
[Eu2(μ2pzdc)(μ4-pzdc)(μ2-ox)(H2O)4]·8H2O 591, 614, 649, 695 [134]
[Eu4(BPT)4(DMF)2(H2O)8]·(DMF)5·(H2O)3 580, 650, 592, 617, 700 [135]
[Eu2(fumarate)2(oxalate)(H2O)4]·4H2O 591, 617, [136]
Eu2(1,4-BDC)3(H2O)2·(H2O)2 590, 617, 698 [75]
49

(continued)
Table 1 (continued)
50

MOF Emission wavelength (nm)Bermudez References


ITQMOF-1-Eu 619 [137]
ITQMOF-3-Eu 579.0, 580.7, 612.3 [138]
Eu2(2,5-pdc)2(1,4-pda)(H2O)2 ~590, 615, ~698 [101]
Na4[Eu2(H2O)2(hedp)2]·H2O ~590, 613 [110]
Eu2(1,4-BDC)3(MeOH)4]·8MeOH ~590, ~615 [139]
[Eu(H2TETA)]NO3·2H2O 579, 592, 613, 651, 698 [125]
Eu2(Hbidc)2(ox)2·(H2O)3 536, 594, 617, 651, 695 [140]
Eu2(ATPA)3(DMF)2(H2O)2 450, 616 [127]
Gd [Gd2(Hpimda)2(μ4-C2O4)·4H2O]·2H2O 319, 476 [122]
[Gd2(pyda)(μ4-C2O4)2·4H2O]·3H2O 299, 546 [122]
Tb Tb(1,4-BDC)3(H2O)4 491, 546 [128]
[Tb(pta)(H2O)5]·4H2O 491, 546, 584, 623 [126]
K5[Tb5(IDC)4(ox)4] 489, 545, 589, 619 [141]
[Tb(Mucicate)1.5·(H2O)2] ·5H2O ~490, ~545 [142]
Tb(BTC)·methanol 492, 548, 584, 620 [143]
TbNTA·H2O 489, 547, 583, 625 [144]
Na[Tb(OBA)2]3·0.4DMF3·1.5H2O 545 [145]
[Tb4(BPT)4(DMF)2(H2O)8]·(DMF)5·(H2O)3 490, 544, 587, 621, 650 [135]
[Tb2(fumarate)2(oxalate)(H2O)4]·4H2O 494, 545, 587, 621 [136]
Tb2(2,5-pdc)2(1,4-pda)(H2O)2 544, ~585, ~620 [101]
Tb(4-hydroxybenzenesulfonate)3(H2O)2·2H2O ~490, ~545 [146]
Tb3(1,4-BDC)4.5(DMF)2(H2O)3·(DMF)(H2O) 370, 430, 490, ~550, ~590 [147]
Tb(1,4-BDC)(MeOH)4] ·Cl·MeOH·0.25H2O ~495, ~550 [139]
[Tb(H2TETA)]NO3·2H2O 492, 545, 584, 620 [125]
[Tb2(Hbidc)2(ox)(H2O)2]·4H2O 490, 548, 590, 620 [140]
Tb2(ATPA)3(DMF)2(H2O)2 450, 489, 545, 587, 619 [127]
Dy Dy(1,4-BDC)3(H2O)4 481, 575 [128]
476, 569 [126]
Y. Cui et al.

[Dy(pta)(H2O)3]·4H2O
K5[Dy5(IDC)4(ox)4] 479, 573 [141]
[Dy4(BPT)4(DMF)2(H2O)8]·(DMF)5·(H2O)3 480, 573, 662 [135]
Dy2(ATPA)3(DMF)2(H2O)2 450 [127]
Er [Er(pta)(H2O)3]·4H2O 1,538 [126]
Er2(1,4-BDC)3(DMF)2(H2O)2·H2O 1,550 [148]
Er2(BDC-F4)3(DMF)(H2O)·DMF 1,550 [148]
Yb [Yb(pta)(H2O)3]·4H2O 980 [126]
Yb(BPT)(H2O)·(DMF)1.5(H2O)1.25 980 [149]
[Yb2(C26H20O6)3(H2O)2]·(DMF)6(H2O)8.5 980 [150]
[Yb2(C26H20O6)3(H2O)2]·(DMF)12(H2O)10 980 [150]
(Yb2(TDC)3(DMF)(H2O)·(CH3CH2OH) 980 [151]
Transition-metal-based MOFs
Zn [Zn4O(NTB)2]·3DEF·EtOH 433 [152]
Zn4O(1,4-BDC)3 397 [153]
Zn4(OH)2(1,2,4-BTC)2 428 [154]
(Zn3(TAA)3(H2O)3·(1,4-dioxane) 438 [155]
[Zn4O(1,4-BDC)(bpz)2]·4DMF·6H2O 470 [156]
Zn2(bpdc)2(bpee) 420 [157]
Zn–BCPA 438 [158]
Zn3(μ5-pta)2(μ2-H2O)2 467 [92]
Zn3(BTC)2(DMF)3(H2O)·(DMF)(H2O) 410 [93]
Zn3(4,40 -stilbene dicarboxylic acid)3(DMF)2 396, 420, 441 [159]
Luminescent Properties and Applications of Metal-Organic Frameworks

[Zn(2,3-pydc)(bpp)]·2.5H2O 436 [112]


Zn2(ATA)3(ATA)2/2 485 [114]
Zn(hfipbb)·0.5H2O·0.5DMF 425 [160]
Zn(hfipbb)(4,40 -bipy)·DMF 425 [160]
Zn5(μ3-OH)2(BTA)2(tp)3 348 [161]
Zn(BTA)(chdc)0.5 380 [161]
Zn(BTA)(ap)0.5 360 [161]
Zn(BTA)(gt)0.5 358 [161]
51

(continued)
Table 1 (continued)
52

MOF Emission wavelength (nm)Bermudez References


Zn(DFDA) 405, 422 [162]
Zn(DFDA)(1,10 -(1,4-butanediyl)bis(imidazole)) 405, 427 [162]
Zn(DFDA)( 2,20 -bipyridine) 411, 427 [162]
Zn2(3-tzba)(N3)(OH)(2,20 -bipy) 420 [163]
[Zn(3-tzba)(2,20 -bipy)(H2O)]·3H2O 437, 465 [163]
[Zn2(3-tzba)2(phen)2]·H2O 427 [163]
Zn(sfdb)(bpy)(H2O)]·0.5CH3OH 359 [164]
Zn3(CTC)2(bipy)·(DMF)(H2O)2 447, 552 [165]
Zn3(CTC)2(Bpe)·(DMF)(H2O)2 456, 560 [165]
Cd Cd(BDC–NH2)(bpy)·4H2O·2.5DMF 435 [166]
Cd4(BTC)3(DMF)2(H2O)2·6H2O 405 [93]
[Cd(2,3-pydc)(bpp)(H2O)]·3H2O 438 [112]
[Cd(dtba)(bpp)]3n 434, 482 [113]
Cd(hfipbb)(DMF)·0.5DMF 425 [160]
Cd(5MT)2 406 [167]
Cd5(N3)(5MT)9·0.12H2O 405 [167]
Cd3(OH)Cl1.39(N3)0.61(5MT)3 412 [167]
Cd3(OH)Cl(N3)(ATA)3 422 [167]
Cd2(OH)Br(ATA)2 421 [167]
Cd7Cl2(5BT)12(H2O)2 394 [167]
Cd7Br2(5BT)12(H2O)2 395 [167]
Cd2(BPVIC)2(SCN)4 562 [168]
Cd(DFDA)(C2H5OH) 418 [162]
Cd2(DFDA)2(bis(imidazol-1-ylmethyl)benzene)2 407, 428 [162]
[Cd(sfdb)(phen)2]·2H2O 380 [164]
[Cd(sfdb)(bpy)2]·H2O 396 [164]
Cd(sfdb)(quin) 385 [164]
394 [164]
Y. Cui et al.

Cd3(sfdb)2(Hsfdb)2(phen)2
Cd(sfdb)(bpy) 365 [164]
Cd2(ABTC)(DMF)3·(DMF)2 402 [169]
Cd3(CTC)2(TED)(H2O)2·(H3O)2Cl2 522 [165]
Cd3(CTC)2(bipy)(DMF)2·(DMF)(H2O)2 449, 550 [165]
Cd3(CTC)2(Bpe)(DMF)2·(DMF)(H2O)2 453, 556 [165]
Ag [Ag(4-cyanobenzoate)L]·H2O 427, 513, 566, 617 [170]
[Ag8(L)4](NO3)8·4H2O, L ¼ bis(3,5-bis((1H- imidazol-1-yl)-methyl)-2,4,6-trimethylphenyl) 495 [171]
methane
[Ag2(L)2(ClO4)2(CH3CN)], L ¼ N,N0 -bis(pyridin-2-ylmethylene)benzene-1,4-diamine 451, 521 [172]
[Ag2(L)2(ClO4)2], L ¼ 3,30 -dimethyl-N,N0 -bis(pyridin-2-ylmethylene)biphenyl-4,40 -diamine 409, 432, 505, 526 [172]
[AgL]ClO4, L ¼ N0 ,N0 -bis[1-(pyridin-4- yl)methylidene]benzil dihydrazone 340 [173]
[AgL2]BF4, L ¼ N0 ,N0 -bis[1-(pyridin-4- yl)methylidene]benzil dihydrazone 342 [173]
[AgL(NO3)](CH3CH2OH), L ¼ N0 ,N0 -bis[1-(pyridin-4- yl)methylidene]benzil dihydrazone 339 [173]
[AgL(NO3)], L ¼ N0 ,N0 -bis[1-(pyridin-3- yl)methylidene]benzil dihydrazone) 344 [173]
[Ag2(PhPPy2)2Cl](ClO4) ~510 [174]
Ag(MES)(L)0.5, L ¼ 1,10 -(1,4- butanediyl)bis(2-methylbenzimidazole) 479 [175]
[Ag(MES)(L)0.5]·H2O, L ¼ 1,10 -(1,4-butanediyl)bis(2-ethylbenzimidazole) 363 [175]
Cu Cu6(5,6-diphenyl-1,2,4-triazine-3-thiol)6·(H2O)(DMSO) 660 [176]
Cu3(L)2·3H2O, L ¼ chelidamic acid 398, 478 [111]
Cu5(SCN)5(3-Abpt)2 559 [177]
Cu(SCN)(3-Abpt) 570 [177]
Cu2(SCN)2(4-PyHBIm) 569 [177]
Luminescent Properties and Applications of Metal-Organic Frameworks

Cu2(SCN)2(3-PyHBIm) 530 [177]


Cu(SCN)(4-Ptz) 435 [177]
Cu2(SCN)2(2-PyBIm)(2-PyHBIm) 440 [177]
[Cu12Br2(CN)6/2(SCH3)6][Cu(SCH3)2] 575 [178]
Cu2I2(DABCO)2 551 [43]
Cu2I2(piperazine)2 560 [43]
Cu2(CN)2(bpzm) 473 [179]
Cu2(CN)2(bpze) 529 [179]
53

(continued)
Table 1 (continued)
54

MOF Emission wavelength (nm)Bermudez References


Cu2(CN)2(bpzp) 515 [179]
[Cu(Pz)]3 542 [180]
[Cu2(Bpz)]n 598 [180]
Cu2(ABTC)(H2O)2·(DMF)2(H2O) 392 [169]
Co Co(L)(2,20 -bipy), L ¼ dibenzothiophene- 5,50 -dioxide-3,7-dicarboxylate 409 [181]
Mn Mn(Hbidc) 726 [182]
[MnL2(NCS)2]·(CH3CH2OH), L ¼ N0 ,N0 -bis[1- (pyridin-3-yl)methylidene]benzil dihydrazone) 356 [173]
Fe [FeL2Cl2]·CH3CN, L ¼ N0 ,N0 -bis[1-(pyridin-3-yl) methylidene]benzil dihydrazone) 353 [173]
Heterometal-organic frameworks
Pr–Cu Pr(pydc)3Cu3(bipy)3·m(H2O) 488, 532 [183]
Nd–Cd [NdCd(imdc)(SO4)(H2O)3]·0.5H2O 430 [184]
Nd–Cu Nd(pydc)3Cu3(bipy)3·5(H2O) 895, 1,062, 1,345 [183]
Sm– [SmAg(PDA)2(H2O)3]·3H2O 563, 601, 644 [185]
Ag
Sm–Cu Sm(pydc)3Cu3(bipy)3·4(H2O) 563, 603, 645 [183]
SmCu(nds)(isonicotinic acid)2·H2O 561, 599, 645 [186]
Sm–Ba Ba2(H2O)4[SmL3(H2O)2](H2O)nCl, L ¼ 4,40 -disulfo-2,20 - bipyridine-N,N0 -dioxide ~605, ~640 [187]
Eu–Mn [Eu(PDA)3Mn1.5(H2O)3]·3.25H2O ~590, 618 [188]
Eu–Fe [Eu(PDA)3Fe1.5(H2O)3]1.5H2O 613 [189]
Eu–Ag [EuAg(PDA)2(H2O)3]·3H2O 581, 593, 617, 651, 695 [185]
EuAg(pydc)2(Hnic)0.5 ~580, ~615 [190]
EuAg(inic)2(nicO)·0.5H2O ~580, ~620 [190]
Eu–Ba Ba2(H2O)4[EuL3(H2O)2](H2O)nCl, L ¼ 4,40 -disulfo-2,20 - bipyridine-N,N0 -dioxide ~615 [187]
Eu–Cd [EuCd(imdc)(SO4)(H2O)3]·0.5H2O ~590, ~618 [184]
Eu–Cu Eu(pydc)3Cu3(bipy)3·4(H2O) 579, 593, 615, 651 [183]
[EuCu(nic)2(ox)]·2H2O 579, 591, 594, 616, 618 [191]
EuCu(nds)(isonicotinic acid)2·H2O 579, 592, 619, 654, 704 [186]
Gd–Cd [GdCd(imdc)(SO4)(H2O)3]·0.5H2O 355, 400–465, 544 [184]
Y. Cui et al.
Gd–Cu Gd(pydc)3Cu3(bipy)3·4(H2O) 488, 532 [183]
Gd–Mn [(Gd2L)Mn(H2O)6]·0.5(H2O), L ¼ 3,30 -(4-Amino-4H-1,2,4-triazole-3,5-diyl)dibenzoic acid 428 [192]
Gd–Ba Ba2(H2O)4[GdL3(H2O)2](H2O)nCl, L ¼ 4,40 -disulfo-2,20 - bipyridine-N,N0 -dioxide 416, 435, 448 [187]
Tb–Mn [Tb(PDA)3Mn1.5(H2O)3]·3.25H2O ~490, ~550 [188]
Tb–Ag [TbAg(PDA)2(H2O)3]·3H2O 490, 545, 584, 622 [185]
Tb–Cd [TbCd(imdc)(SO4)(H2O)3]·0.5H2O 490, 544, 584, 620 [184]
Tb–Cu Tb(pydc)3Cu3(bipy)3·4(H2O) 492, 544, 583, 622 [183]
TbCu(nds)(isonicotinic acid)2·H2O 489, 544, 590, 619 [186]
Tb–Ba Ba2(H2O)4[TbL3(H2O)2](H2O)nCl, L ¼ 4,40 -disulfo-2,20 - bipyridine-N,N0 -dioxide ~490, ~545 [187]
Dy–Ag [DyAg(PDA)2(H2O)3]·3H2O 483, 573 [185]
Dy–Cd [DyCd(imdc)(SO4)(H2O)3]·0.5H2O 478, 573, 657 [184]
Dy–Ba Ba2(H2O)4[DyL3(H2O)2](H2O)nCl, L ¼ 4,40 -disulfo-2,20 - bipyridine-N,N0 -dioxide ~640 [187]
Ho–Ag HoAg5(1,2-BDC)4 995, 1,400–1,600 [193]
Er–Ag ErAg5(1,2-BDC)4 1,450–1,650 [193]
ErAg3(L)3(H2O), L ¼ pyridine-2,6-dicarboxylic acid 1,548 [194]
ErAg3(L)2(H2O), L ¼ 4-hydroxylpyridine- 2,6-dicarboxylic acid 1,540 [194]
Er–Cu Er(pydc)3Cu3(bipy)3·4(H2O) 488, 532 [183]
Yb–Cd [YbCd(imdc)(SO4)(H2O)3]·0.5H2O 433 [184]
Yb–Ag YbAg5(1,2-BDC)4 950–1,050 [193]
Yb–Cu Yb(pydc)3Cu3(bipy)3·4(H2O) 980 [183]
Eu–Ce [Eu(dipicH)(H2O)6][Ce(dipic)3]3·7H2O ~590, ~615 [195]
Eu–Tb Eu1-xTbx(BTC)(H2O) 540, 589, 615 [196]
Luminescent Properties and Applications of Metal-Organic Frameworks

Tb(1,3,5-BTC)(H2O)·3H2O:Eu3+ 487, 543, 580, 620 [197]


[(Eu, Tb)(C6H8O4)3(H2O)2]·(C10H8N2) 545, 595, 615 [100]
(Eu0.2Tb0.8)2(2,5-pdc)2(1,4-pda)(H2O)2 ~545, ~595, ~620 [101]
Eu0.0069Tb0.9931-DMBDC 489, 545, 592, 613 [198]
Dy–Ce [Dy(dipicH)(H2O)6][Ce(dipic)3]3·7H2O ~480, ~575 [195]
Er–Yb ErxYb1-x-PVDC-1 980, 1,530 [199]
Zn–Ir [Zn4(μ4-O)(Ir(2-pyridyl-benzoic acid)3)2]·6DMF·H2O 538 [200]
(continued)
55
Table 1 (continued)
56

MOF Emission wavelength (nm)Bermudez References


Cu–Ag CuAg2(L)2, L ¼ chelidamic acid 515 [111]
Others
In In2(OH)2(TBAPy) 471 [201]
[In3(BTB)2(ox)3]n ~460 [202]
Bi Bi3(μ3-O)2(pydc)2(Hpydc)(H2O)2 430, 460, 480, 556 [203]
Pb Pb(pydc)(H2O) 441, 470, 520, 563 [203]
Pb4(1,3-BDC)3(μ4-O)(H2O) 424 [204]
Pb(bpdc) 480 [205]
Y. Cui et al.
Luminescent Properties and Applications of Metal-Organic Frameworks 57

the lighting, display, and optical devices. The robust pores within luminescent
MOFs are particularly useful to develop luminescent sensing materials. In fact, to
make use of the pores within luminescent MOFs for their differential recognition of
sensing substrates, a number of porous luminescent sensing MOFs have been
fulfilled. Luminescent MOFs might also be utilized as light-emitting materials for
white-light and near-infrared light-emitting devices. Nanoscale luminescent MOFs
have started to be explored for their applications on tissue and cell imaging, as well
as drug delivery monitoring and treatment.

4.1 Near-Infrared (NIR) Luminescent Materials

Recent researches in the field of lanthanide-based near-infrared (NIR) luminescent


materials are very active due to their advantages for lasers, optical amplifiers for
telecommunication, biological applications, NIR organic light-emitting diode
(OLED) technology, solar energy conversion, etc. For example, lanthanide ions
Nd3+ and Er3+ exhibit near-infrared luminescence at wavelengths of 1,330 and
1,550 nm, respectively, which are of interest for applications in optical telecommuni-
cation networks. However, a few examples of NIR light-emitting lanthanide
frameworks are reported in comparison to visible-emitting MOFs due to their smaller
energy gap between the ground state and emissive state of these ions, as compared to
Eu3+ and Tb3+, thus resulting in more efficient nonradiative decay through high-
energy C–H, N–H, and O–H oscillators in the ligands and solvents [96].
In order to alleviate nonradiative decays, a variety of strategies have been
adopted to shield the ion excited levels to high nonradiative transition probability
by O–H, C–H, and N–H oscillators. Chen et al. demonstrated that the NIR lumines-
cence intensity of MOFs can be significantly enhanced by incorporation of
fluorinated organic linkers. They synthesized two erbium–organic frameworks
Er2(1,4-BDC)3(DMF)2(H2O)2·H2O (1,4-BDC ¼ 1,4-benzenedicarboxylate) and
Er2(BDC–F4)3(DMF)(H2O)·DMF (BDC–F4 ¼ 2,3,5,6-tetrafluoro-1,4-benzenedi-
carboxylate) [148]. After heat treatment under vacuum at 140 C overnight, the
desolvated framework displays NIR emission band around 1.55 μm assigned to the
4
I13/2 ! 4I15/2 transition of Er3+ ion at the excitation wavelength of 808 nm.
Interestingly, the emission intensity of Er–BDC–F4 framework is significantly
enhanced by 3 times than that of Er–BDC framework because fluorination can
significantly reduce the fluorescence quenching effect of the vibrational C–H bond.
Chen et al. speculated that the luminescent intensity of the fluorinated framework
can be further enhanced by incorporating ancillary electron-rich organic ligands
instead of solvent molecules to saturate the Er3+ coordination sites.
Additionally, a series of NIR luminescent Ln–Ag HMOFs, LnAg5(1,2-BDC)4
(Ln ¼ Yb, Er, Ho, 1,2-BDC ¼ 1,2-benzenedicarbolylate) have been reported by Jin
et al. [193]. The HMOF YbAg5(1,2-BDC)4 shows a broad emission band ranging
from 950 to 1,050 nm assigned to the 2F5/2 ! 2F7/2 transition of the Yb3+ ion, and the
ErAg5(1,2-BDC)4 exhibits a broad emission band extending from 1,450 to 1,650 nm
58 Y. Cui et al.

attributed to the 4I13/2 ! 4I15/2 transition of an Er3+ ion. The HoAg5(1,2-BDC)4


displays an emission at 995 nm assigned to the 5F5 ! 5I7 transition of Ho3+ ion and a
band from 1,400 to 1,600 nm corresponding to the splitting of 5F5 ! 5I6 transition.
Especially, the NIR emission for these HMOFs show shift, splitting or broadening
emission spectra compared with the isolated lanthanide ions, which might be
attributed to the interaction and influence of 4d and 4f orbitals. Jin et al. speculated
´
that the Ag–Ln distance becomes much shorter (ranging from 3.71 to 3.75 Å) upon
formation of the Ag–Ln HMOFs, so the 4d orbital of the Ag+ ion and 4f orbital of the
Ln3+ ions may interact and influence each other, resulting in the tuned inner levels of
the Ag–Ln system and shifting and/or split of the lower-energy states of Ln3+ ions.
Other NIR luminescent HMOFs include an infinite triple-stranded helical frame-
work ErAg3(L)3(H2O) (L ¼ pyridine-2,6-dicarboxylic acid) and a sandwich-type
3D framework ErAg3(HO–L)2(H2O) (HO–L ¼ 4-hydroxylpyridine-2,6-dicarbox-
ylic acid) [194]. The HMOF ErAg3(L)3(H2O) shows a broad emission band at
1,548 nm ascribed to the 4I13/2 ! 4I15/2 transition of Er3+ ion. Similarly,
ErAg3(HO–L)2(H2O) exhibits an emission band with the strongest peak at 1,540 nm.
Several examples of Ln–Cu HMOFs with mixed bridging ligands Ln
(pydc)3Cu3(bipy)3·m(H2O) (Ln ¼ Pr, Nd, m ¼ 5; Ln ¼ Sm, Eu, Gd, Tb, Er, Yb,
m ¼ 4; pydc ¼ 2,6-pyridinedicarboxylate anion; bipy ¼ 4,40 -bipyridine) were
reported by Bo et al. [183]. During the self-assembly process, the heterometallic ions
are interconnected by mixed ligands to produce spindle-shaped heterometallic rings
[Ln6(pydc)6Cu12(bipy)6] as the SBUs which are pillared by bridging bipy molecules to
yield the porous 3D pillared-layer Ln–Cu HMOFs. The Nd–Cu HMOFs display the
strongest emission band at 1,062 nm ascribed to 4F3/2 ! 4I11/2 transition of Nd3+ ion, a
broad emission band centered at 895 nm attributed to 4F3/2 ! 4I9/2 transition, and a
sharp weak emission at 1,345 nm corresponding to 4F3/2 ! 4I13/2 emission. The Yb–Cu
HMOF also displays a strong emission at 980 nm attributed to the 2F5/2 ! 2F7/2 transi-
tion of the Yb3+ ion in near-infrared region upon the excitation at 275 nm. All Sm–Cu,
Eu–Cu, and Tb–Cu HMOFs exhibit the characteristic emission of the corresponding
lanthanide ions; while the Pr–Cu, Gd–Cu, and Er–Cu HMOFs display two intense
emission peaks at 488 and 532 nm, similar to those found in Cu(I) complexes, which
are assigned to the emission from MLCT transition.
Although some excellent ligands with high intersystem crossing yields such as
β-diketonates and triphenylene derivative have been designed for sensitizing lantha-
nide ions; a drawback of these sensitizers is that they require near-UV excitation and
that they have relatively high-lying triplet energy levels. White et al. utilized a new
ligand to efficiently sensitize the NIR emission in the Yb3+-MOFs. Based on the
ligand 4,40 -[(2,5-dimethoxy-1,4-phenylene)di-2,1-ethenediyl]bisbenzoic acid (H2-
PVDC), two MOFs, named [Yb2(PVDC)3(H2O)2]·(DMF)6(H2O)8.5 (Yb-PVDC-1)
and [Yb2(PVDC)3(H2O)2]·(DMF)12(H2O)10 (Yb-PVDC-2) were synthesized
[150]. These MOFs display the NIR emission at 980 nm of Yb3+ ion, but the
excitation band of Yb-PVDC-2 is red shifted to 500 nm from 470 nm in
Yb-PVDC-1, which was proposed that the close π–π interactions between the
PVDC linkers decrease the energy of the π ! π* transition, resulting in a lowered
excitation energy. Additionally, the quantum yield of Yb-PVDC-2 is five times
Luminescent Properties and Applications of Metal-Organic Frameworks 59

Fig. 4 (a) Crystal structure of Yb-PVDC-1 viewed along the crystallographic c axis. (b) PXRD
patterns for Yb-PVDC-1 and analogs ErxYb1-x-PVDC-1 (x ¼ 0.32, 0.58, 0.70, and 0.81). (c) Yb3+
(980 nm) and Er3+ (1,530 nm) emission spectra of ErxYb1-x-PVDC-1 (x ¼ 0.32, 0.58, 0.70, and
0.81) normalized to the Er3+ signal upon 490-nm excitation. (d) Plot of the ratio of integrated
emission intensities versus their atomic ratio. (e) Color-coded schematic of the barcode readout.
Reprinted with the permission from [199]. Copyright 2009, American Chemical Society

higher than that of Yb-PVDC-1 when excited through the lower energy band
(490 nm). The quantum yield is among the highest values reported for Yb3+ systems
under solvent, which may be attributed to the different coordination environments
and levels of protection in both MOFs. This example demonstrates a new route for
controlling the luminescent properties of lanthanide MOFs without altering the
structure of the sensitizer.
Furthermore, White et al. proposed a new approach for creating luminescent
barcoded systems by doping multiple Ln3+ ions into the abovementioned MOFs,
which simultaneously emit several independent NIR signals arising from different
lanthanide compositions [199]. They designed ErxYb1-x-PVDC-1 (x ¼ 0.32, 0.58,
0.70, and 0.81) MOFs with different amounts of Yb3+ and Er3+ ions that were
sensitized by the same ligand PVDC. These MOFs displayed sharp signals from
both Er3+ (1,530 nm) and Yb3+ (980 nm), and the excitation spectrum of either the
Er3+ or Yb3+ emission band contains two similar bands with apparent maxima at
370 and 470 nm. By controlling reactant stoichiometry, the lanthanide composition
of the MOF and the resulting individual emission intensities can be tuned and
controlled (Fig. 4). This features their potential usage as NIR barcodes. Further-
more, the number and diversity of barcodes may be increased by using a large
60 Y. Cui et al.

number of Ln/Ln ratios or by incorporating additional lanthanide ions into the


material. The later concept was demonstrated in MOF Nd0.09Er0.55Yb0.36-PVDC-1,
which displays a more sophisticated barcode signal consisting of NIR emissions
from three lanthanide ions. These MOFs retain their NIR luminescence when
coated in superglue, demonstrating the possibility for practical application.

4.2 White-Light-Emitting Materials

White-light-emitting materials and devices have attracted significant attention due to


their broad applications in displays and lighting. High-quality white-light illumination
requires a source with the Commission International ed0 Eclairage (CIE) coordinates
(0.333; 0.333), with correlated color temperature (CCT) between 2,500 and 6,500 K,
and color rendering index (CRI) above 80 [206, 207]. Emission from organic or
inorganic luminescent materials can only cover part of the visible spectrum. To
overcome this limitation, various architectures of devices combining monochromatic
emission from different compounds have been suggested. In principle, white-light-
emitting devices may be obtained by tuning the emission color and controlling the
relative amount of these monochromatic emissions. In luminescent MOFs, besides the
characteristic f–f emission of lanthanide ions, the broad emission band ascribed to
organic linkers can be observed. The relative luminescence intensity between the
lanthanide ions and linkers strongly depends on the nature of the energy transfer,
thus allowing the fine-tuning of the MOFs emission color across the CIE diagram.
In addition, the emission color can also be readily modulated both by chemical factors
(Ln3+ types and concentration, ligand structure, coordination status, guest species) and
physical parameters (excitation wavelength and temperature).
Doping lanthanide ions into the isostructural MOF materials has been
demonstrated as an effective approach to tune the emission colors. Qian and
Chen et al. reported a new two-dimensional lanthanide MOF La2(PDA)3(H2O)5
[121], which exhibited a blue emission at 408 nm attributed to the emissive organic
PDA linkers. Compared with luminescence of the free ligand, La2(PDA)3(H2O)5
exhibits much enhanced blue light emission and slight red shifts, which was
attributed to the enhancement of rigidity of the PDA ligands after coordinating to
La3+ ions. Doping of a small amount of Tb3+ and Eu3+ led to the formation of
isostructural La2(PDA)3(H2O)5:Tb3+ and La2(PDA)3(H2O)5:Eu3+, which can pro-
vide additional green (543 nm) and red (614 nm) emissions, respectively. There-
fore, the white-light-emitting MOFs La2(PDA)3(H2O)5:Tb3+, Eu3+ can be realized
by tuning the molar amount of Tb3+ and Eu3+ in the host framework solid. The
optimized white-light-emitting MOFs, La2(PDA)3(H2O)5:1.0%Tb3+, 2.0%Eu3+ and
La2(PDA)3(H2O)5:1.5%Tb3+, 2.0%Eu3+, exhibit the CIE coordinate of (0.3269,
0.3123) and (0.3109,0.3332), respectively.
Dang et al. prepared a family of lanthanide MOFs using a semirigid trivalent
carboxylic acid as linker [208]. These MOFs are isostructural and display typical
lanthanide luminescent emissions of Eu3+, Tb3+, Sm3+, and Dy3+, respectively. The
absence of obvious emission from the ligand implies the efficient energy transfer
Luminescent Properties and Applications of Metal-Organic Frameworks 61

from the semirigid ligands to Ln3+ ions. Furthermore, white-light emission was
successfully obtained by codoping Dy/Eu and Dy/Sm into the Gd analogs. The CIE
coordinates of Dy0.02Eu0.05Gd0.93L are (0.355, 0.313), which is very close to the
ideal coordinate for pure white light (0.333, 0.333). Another sample by codoping
Dy and Sm into Gd analog, Dy0.01Sm0.10Gd0.89L, also exhibits the white-light
emission with a CIE coordinates of (0.328, 0.320).
Recently, the pursuing of single white-light emitting materials to avoid the
intrinsic color balance and device complication when using multiphosphors or
multi-LEDs has attracted much attention. A direct white-light-emitting MOF,
[AgL]n·nH2O (L ¼ 4-cyanobenzoate), with tunable yellow to white luminescence
by variation of excitation light was reported by Wang et al. [170]. The MOF displays
a maximum emission at around 427 and 566 nm when excited at 355 and 330 nm,
respectively. When irradiated at 330 nm, the emission intensity at 427 nm is
decreased, but the emission intensity at approximately 513, 566, and 617 nm are
enhanced, generating yellow luminescence. When adjusting the excitation light to
350 nm, the emission peaks at 427 and 566 nm are very strong which results in direct
white light to the naked eye (Fig. 5). The CIE chromaticity coordinates of the white-
light emissions excited at 350 and 349 nm are approximately (0.31, 0.33) and (0.33,
0.34), respectively, comparable to that of pure white light. Therefore, the MOF can
be tuned from yellow to white by direct variation of excitation wavelength,
indicating it can be potentially used as a single white phosphor for a white-light-
emitting device equipped with a deep UV GaN LED, which has light output at
325–350 nm.
Wibowo et al. also presented two single component white-light-emitting
phosphors based on the bismuth and lead MOFs [203]. When excited at 380 nm, the
bismuth MOF Bi3(μ3-O)2(pydc)2(Hpydc)(H2O)2 emits white composite light with a
broad emission spectrum from about 400–600 nm, and three distinct maxima at
430, 460, and 480 nm with a shoulder around 556 nm can be identified, which is
significantly blue shifted compared to the emission of ligand. This blue shift is
attributed to LMCT transition and a change in the intraligand π ! π* and/or
n ! π* transitions. Similarly, the lead MOF Pb(pydc)(H2O) exhibits slightly
“whiter” composite photoluminescence, with a distinct emission maximum at
441 nm and three broad shoulders around 470, 520, and 563 nm. These studies have
paved the way for the utilization of luminescent MOFs in white-light emitting
devices.
Nenoff et al. reported the intrinsic broadband direct white-light emission
originating from a single component MOF [202], which was synthesized based on
indium ions, H3BTB, and ox (BTB ¼ 1,3,5-tris(4-carboxyphenyl)benzene, ox ¼
oxalic acid), namely, [In3(BTB)2(ox)3]n. Interestingly, the material already emitted a
white light owing to the broadband emission over the entire visible light region.
In order to improve the intrinsic color properties such as color rendering index (CRI),
correlated color temperature (CCT), and chromaticity to approach the requirements
for solid-state lighting, the authors introduced a narrowband, red emission compo-
nent into this system via Eu3+ doping. The simple mixing of Eu3+ metal source into
the reaction solution of the indium framework successfully led to the formation of a
62 Y. Cui et al.

Ag
O
N
C
H
b
10.832(3)
c 6.650(3)

b
λex
6.5x104
355
6.0x104 352
5.5x104 351
350
5.0x104 33
0
349 0
35

4.5x10 4 345
330
4.0x104
Counts / cps

3.5x104
3.0x104
2.5x104
2.0x104
427
1.5x104 Noise
4
1.0x10
5.0x104 566
0.0
400 450 500 550 600 650 700 750
λ / nm

Fig. 5 (a) 3-D packing diagram of [Ag(4-cyanobenzoate)]n·nH2O viewed along the a direction.
(b) Solid-state PL spectra of [Ag(4-cyanobenzoate)]n·nH2O by variation of excitation light under the
same metrical condition. Inset shows the PL images of a sample excited by 350- and 330-nm light,
respectively. Reprinted with the permission from [170]. Copyright 2009, American Chemical Society

mixed-metal framework. By increasing the Eu3+ concentration to 10%, the CRI and
CCT shifted closer to the set target of CRI 90 and CCT 3,200 K. When the sample
is excited at 350, 360, 380, and 394 nm, the coordinates are (0.369, 0.301), (0.309,
0.298), (0.285, 0.309), and (0.304, 0.343), respectively, which closely approach
targeted values (Fig. 6). The absolute quantum yield of the mixed MOF was found
to be 4.3% when excited at 330 nm, which is modest as compared to traditional
phosphors, but falls within the expected range for similar reported systems. This
Luminescent Properties and Applications of Metal-Organic Frameworks 63

Fig. 6 Molecular building blocks in MOF [In3(BTB)2(ox)3]n, corresponding to (a) two topologically
distinct 3-connected nodes and (b) a 4-connected node; (c) corner-sharing tetrahedra forming a
6-membered ring; and (d) topological representation of single net in [In3(BTB)2(ox)3]n; (e) emission
spectra of 10% Eu-doped [In3(BTB)2(ox)3]n when excited between 330 and 380 nm; inset: 1931 CIE
chromaticity diagram highlighting corresponding chromaticity coordinates (A–D) approaching
targeted values. Reprinted with the permission from [202]. Copyright 2012, American Chemical
Society

result provides a new path for the rational design of alternative materials for white-
light emission.

4.3 Chemical Sensors

Of all the MOF applications, chemical sensor is clearly a promising application due to
the luminescent properties of MOFs are very sensitive to and dependent on their
structural characteristics, coordination environment of metal ions, nature of the pore
surfaces, and their interactions with guest species through coordination bonds and π–π
interactions and hydrogen bonding. By the incorporation of analyte into their pores, the
MOFs can exhibit different degrees of luminescence enhancement or quenching
behaviors, which make the MOF materials ideal for molecular recognition. The
permanent porosity within some porous luminescent MOFs has enabled the reversible
uptake and release of some sensing substrates; thus the exploration of reversible
luminescent sensing MOFs will be feasible, so sensing MOFs can be regenerated
and repeatedly utilized. The high surface areas make the MOFs serve as a
preconcentrator to concentrate analytes to levels high above those in the surrounding
atmosphere, thus leading to much smaller detection levels than otherwise available.
Moreover, the tunable pore sizes and functional sites such as Lewis basic/acidic sites
and open metal sites afford unprecedented selectivity either through size exclusion or
chemical interactions, while mesoporous nature of some mesoporous luminescent
MOFs will make sensing of some large molecules such as biologically active species
64 Y. Cui et al.

possible. Over the past few years, a wide range of luminescent MOFs for sensing
cations, anions, small molecules, vapors, and explosives have been realized and
reported. In these MOF sensors, lanthanide MOFs are more attractive candidates
than transition-metal MOFs, because many lanthanide MOFs have sharp and charac-
teristic emissions, which can be modulated by both analyte–metal and analyte–ligand
interactions since the unique luminescence mechanism of lanthanide ions.

4.3.1 Cation and Anion Sensor

Cu2+ is one of the most essential and important ions in living biological systems,
particularly in the brain. The efficient, straightforward and real-time detection of
trace amounts of Cu2+ in the brain is helpful to diagnose and treat the diseases
induced by copper metabolism disorders [209, 210]. By immobilizing the Lewis
basic sites within porous MOFs as sensing sites, Chen et al. develop a new way of
sensing metal ions. These types of MOFs are generally difficult to prepare because
such Lewis basic sites tend to bind other metal ions to form condensed structures.
Exploiting the preferential binding of lanthanide ions to carboxylate oxygen atoms
over pyridyl nitrogen atoms in Ln3+–pyridinecarboxylate MOF, they synthesized a
luminescent MOF Eu(PDC)1.5(DMF)·(DMF)0.5(H2O)0.5 (PDC ¼ pyridine-3,
5-dicarboxylate) [131]. The MOF is porous, and the basic Lewis sites line the interior
of the channels, which may be emptied of DMF and water molecules by heating at
200 C. Due to the different binding of these free Lewis basic pyridyl sites with metal
ions, the activated MOF Eu(PDC)1.5 exhibits a sensing function for metal ions: alkali
metal ions and alkaline-earth metal ions have a negligible effect on the luminescence
intensity, whereas Co2+, especially Cu2+, can reduce the luminescence intensity
significantly (Fig. 7). It is speculated that the binding of the pyridyl nitrogen atoms
to Cu2+ or Co2+ reduces the antenna efficiency of the PDC organic linkers to magnify
the f–f transitions of Eu3+, resulting in quenching of the luminescence. These results
indicate that the Lewis basic sites within porous MOFs are expected to play very
important roles for their recognition of small Lewis acidic molecules and metal ions.
Qian and Chen et al. also demonstrated the first example of Cu2+ sensing in
aqueous solution and simulated physiological aqueous solution using lanthanide
MOFs [132]. They prepared a luminescent MOF named as Eu2(FMA)2(OX)
(H2O)4·4H2O (FMA ¼ fumarate; ox ¼ oxalate), whose luminescent intensity is
heavily metal ion dependent: alkaline metal ion and alkaline-earth metal ions have
basically no effect on the luminescence intensity, while Cu2+ has the largest
quenching effect, and the fluorescence lifetime of 394.60 ms in Eu2(FMA)2(OX)
(H2O)4·4H2O is very significantly reduced to 30.45 ms in the presence of 102 M
Cu2+. The sensing of Cu2+ in the simulated physiological aqueous solution is also
explored, and the comparable result is observed, indicating this MOF is suitable to
probe Cu2+ ion in a biological system. A similar example for sensing of Cu2+ and
Co2+ in aqueous solution is reported by Luo et al. [116]. Instead of lanthanide
MOFs, they utilized a lanthanide-doped approach to obtain luminescent MOF. The
Eu3+- and Tb3+-doped MOFs were prepared by immersing the [NH4]2[ZnL]·6H2O
Luminescent Properties and Applications of Metal-Organic Frameworks 65

Fig. 7 (a) Crystal structure of Eu(PDC)1.5(DMF)(DMF)0.5(H2O)0.5, viewed along a axis


indicating immobilized Lewis basic pyridyl sites oriented toward pore centers. C gray, N purple,
O red, Eu green polyhedra. Hydrogen atoms, terminal DMF molecules, and solvent molecules are
omitted for clarity. (b) Comparison of the luminescence intensity of dehydrated MOF Eu(PDC)1.5
incorporating different metal ions in 10-mM DMF solutions of M(NO3)x. Reprinted with the
permission from [131], Copyright 2009, John Wiley & Sons Ltd.

(L ¼ 1,2,4,5-benzenetetracarboxylate) in EuCl3 or Tb(ClO4)3 solutions through


ion exchange of NH4+. Subsequently, the obtained lanthanide-doped MOFs were
immersed in the MClx (M ¼ Na+, K+, Zn2+, Ni2+, Mn2+, Co2+, Cu2+) solutions to
monitor the luminescent intensity. These doped MOFs show a negligible effect on
the luminescence intensity for the Na+, K+, and Zn2+, while they are highly
sensitive to the Co2+ (in Tb@MOF) and Cu2+ (in Eu@MOF), respectively.
The sensing of anions is also an interesting theme because many inorganic anions
play fundamental roles in environmental and biological systems. Chen et al. reported
a novel luminescent MOF, Tb(BTC)·G (MOF-76, BTC ¼ benzene-1,3,5-
tricarboxylate, G ¼ guest solvent) with OH groups in the terminal solvents instead
of in the organic linkers [143]. MOF-76 exhibits increased luminescence upon
incorporated with Br, Cl, F, SO42, and CO32 in methanol. Because F has
much stronger interactions with the terminal methanol molecules to confine the O–H
bond stretching and thus reduce its quenching effect, the addition of F- enhances the
luminescence intensity by as much as a factor of four over the non-F-incorporated
MOF-76 (Fig. 8). This selective sensing function with respect to F is also observed
in the solvent of DMF, featuring the bright promise for these kinds of porous
luminescent MOFs for the sensing of fluoride anion.
In most examples presented above, the sensing function is mainly based on
lanthanide MOFs; a particular work which relies on ligand-centered fluorescence for
sensing of anions is provided by Qiu et al. [211]. In their report, a highly symmetric
tetragonal cadmium-based MOF with 5-methyl-1H-tetrazole as ligand was
synthesized. In DMF suspension, this MOF exhibits an emission at around 370 nm
attributed to the intraligand transitions, whose intensity decreased continuously upon
adding NO2-, while there was only a negligible effect on the luminescence intensity for
66 Y. Cui et al.

Fig. 8 (a) Single crystal X-ray structure of MOF-76b activated in methanol containing NaF with the
model of fluoride (green) at the center of the channel involving its hydrogen-bonding interaction with
terminal methanol molecules (methanol oxygen, purple; the methyl group from methanol is omitted
for clarity). (b) Excitation (dotted) and PL spectra (solid) of MOF-76b solid activated in different
concentrations of NaF methanol solution (excited and monitored at 353 and 548 nm, respectively).
Reprinted with the permission from [143]. Copyright 2008, American Chemical Society

ClO4, NO3, Cl, and I- and the same luminescence response was also observed in a
water suspension.

4.3.2 Molecule Sensor

Recently, Chen et al. demonstrated that the open luminescent lanthanide sites such
as Eu3+ and Tb3+ will play a crucial role in molecular recognition processes as
binding of substrates [133]. They prepared a MOF Eu(BTC)(H2O)·1.5H2O (BTC ¼
benzenetricarboxylate), which is isostructural with MOF-76 and exists with 1D
channels of about 6.6  6.6 Å (Fig. 9). The luminescence intensity of activated
EuBTC is largely dependent on the solvent molecules, particularly in the case of
DMF and acetone, which exhibit the most significant enhancing and quenching
effects, respectively. Furthermore, a gradual decrease of the fluorescence intensity
was observed upon the addition of acetone to the 1-propanol emulsion of EuBTC,
and the 5D0 ! 7F2 emission intensity of Eu3+ versus the volume ratio of acetone
could be well fitted with a first-order exponential decay. It is suggested that the
binding interaction of the open metal sites with guest solvent molecules definitely
plays an important role, and the weakly coordinated 1-propanol molecules on the
Eu3+ sites can be gradually replaced by DMF and acetone molecules, leading to the
luminescence enhancement and diminishment, respectively.
In addition, Chen et al. prepared a NIR luminescent ytterbium MOF Yb(BPT)
(H2O)·(DMF)1.5(H2O)1.25 (BPT ¼ biphenyl-3, 40 , 5-tricarboxylate) for sensing of
small molecules [149]. This MOF crystallizes in a tetragonal space group P43. Each
ytterbium atom is coordinated by six oxygen atoms from the carboxylate groups of
Luminescent Properties and Applications of Metal-Organic Frameworks 67

a b
2500

D0→ F2
7
2000

5
1500

D0→7F1

D0→7F4
1000

D0→ F3
5

7
500

5
5
0

)
l.%
0.3125

vo
0.625

t(
0.9375

en
1.25

nt
Co
1.875

e
2.5

on
3.75

et
Ac
500 550 600 650 700 750
Wavelength (nm)

Fig. 9 (a) X-ray crystal structure of Eu(BTC)(H2O)·1.5H2O viewed along the c axis, exhibiting
1D channels of about 6.6  6.6 Å, and uniformly immobilized accessible Eu3+ sites within the
framework, shown by the pink arrows. The free and terminal water molecules are omitted for
clarity; Eu, light green; O, red; C, gray; H, white. (b) The PL spectra of 1-propanol emulsion of
EuBTC in the presence of various content of acetone solvent. Reprinted with the permission from
[133], Copyright 2007, John Wiley & Sons Ltd

BPT and one terminal water molecule. Yb atoms are bridged by BPT organic
linkers to form a 3D rod-packing structure. The activated MOF Yb(BPT) exhibits
typical NIR emission from the 2F5/2 ! 2F7/2 transition of the Yb3+ ion at 980 nm,
when excited at 326 nm. Similar to the EuBTC, the luminescent intensity exhibit
the most significant enhancing and quenching effects in the case of DMF and
acetone, respectively. The decreasing trend of the fluorescence intensity of the
2
F5/2 ! 2F7/2 transition of Yb3+ at 980 nm versus the volume ratio of acetone could
be well fitted with a first-order exponential decay, indicating that fluorescence
quenching of Yb(BPT) by acetone is diffusion controlled. This work extends the
sensing function of luminescent MOFs into the NIR region, providing their
potentials for the sensing of substrates in biological systems.
Another example of recognition of small solvent molecules was reported by
Lin et al. [145], who synthesized a new family ytterbium–organic frameworks with
4,40 -oxybis(benzoate) (OBA) ligands and suitable cationic species. The MOF Na
[Tb(OBA)2]3·0.4DMF3·1.5H2O exhibits the strongest emission intensities in BuOH
and EtOH suspensions and significantly decrease in MeOH and H2O suspensions
gradually increasing luminescence intensities when dispersed sequentially in water,
methanol, and ethanol as suspensions. Unlike the sensing result of Eu(BTC)
mentioned above, this MOF does not show its strongest emission in a DMF
suspension, which was suspected that the different sizes and geometries of active
metal centers were exposed in these frameworks, such as six- and eight-coordinated
metal sites in the desolvated Eu(BTC) and Na[Tb(OBA)]2 solids, respectively, thus
resulting in their own unique sensing abilities.
An example for recognition of small solvent molecules based on ligand-centered
luminescence was performed in an In–pyrenetetrabenzoic acid MOF, reported by
Stylianou et al. [201]. They selected a highly fluorescent pyrene-derived ligand with
68 Y. Cui et al.

an indium ion to form a permanently microporous fluorescent metal-organic frame-


work, In2(OH)2(TBAPy) (TBAPy ¼ 1,3,6,8-tetrakis( p-benzoic acid)pyrene). This
MOF shows a strong linker-centered fluorescence emission band at 471 nm and a
longer lifetime about 0.11 ms than that of the uncoordinated TBAPy ligand.
When the desolvated MOF was immersed in different solvents, solvent-dependent
luminescent spectra were observed.
In contrast to the bulky macroscopic MOFs, the excellent dispersible nature of the
nanoscale MOFs enables them to directly interact with chemical species and does not
require to be activated. Qian and Chen et al. reported a rare luminescent nanoscale
MOF Eu2(BDC)3(H2O)2·(H2O)2 (BDC ¼ benzene-1,4-dicarboxylate), for the
straightforward and highly sensitive sensing of nitroaromatic explosives in ethanol
solution [75]. Interestingly, the analytes such as benzene, toluene, chlorobenzene,
phenol, o-cresol, and 4-bromophenol basically do not affect the luminescence inten-
sity, while the nitroaromatic compounds such as nitrobenzene, 2,4-DNT
(2,4-dinitrotoluene), and TNT (2,4,6-trinitrotoluene) significantly quench the lumi-
nescence of Eu2(BDC)3(H2O)2·(H2O)2 in ethanol, which is attributed to a competition
of absorption of the light source energy and the electronic interaction between the
nitroaromatic compounds and BDC ligands. The variety of the organic ligands which
can be assembled into the nanoscale MOFs for their matching competitive absorption
with the different analytes highlights this strategy as a very promising approach for
the sensing of small molecules.
Detection of Bacillus anthracis is an essential process in the event of biological
warfare since these bacterial endospores serve as a delivery vehicle in anthrax attacks.
The luminescent detection of bacterial endospores is based on the presence of DPA
(DPA ¼ dipicolinic acid) in the endospore casing, which weighs up to 15% of the
endospores. Qian and Chen described the highly sensitive, selective, and instant turn-
on sensing of bacterial endospores using a nanoscale MOF, Eu2(FMA)2(OX)
(H2O)4·4H2O (FMA ¼ fumarate, ox ¼ oxalate) [212]. Interestingly, the sizes and
morphologies of the nanoscale MOF can be simply tuned and controlled by the
addition of different amounts of the CTAB surfactant (CTAB ¼ cetyltrimethyl-
ammonium bromide), matching with the simulatation based on the well-established
Bravais–Friedel–Donnay–Harker (BFDH) method very well. The nanoscale nature of
Eu2(FMA)2(OX)(H2O)4·4H2O has not only allowed the easily disperse in solvent, but
also significantly enhanced the sensitivity level up to about 90 times for 2 ppm DPA
sensing. Such highly sensitive sensing of DPA is basically not affected by other
components such as isophthalic acid, phthalic acid, terephthalic acid, 1,3,5-
benzenecarboxylic acid, D-phenylalanine, sodium benzoate, riboflavin, DL-tryptophan,
and different ions which coexist with DPA in the bacterial endospores. This result
suggests very bright promise for developing controllable nanoscale functional MOF
sensors.
The luminescent MOF sensors mentioned above generally detect a single target
molecule among several molecules, but cannot differentiate multiple molecules
simultaneously. Recently, a clever molecular decoding strategy using MOFs as
hosts has been developed by Takashima et al. [213]. Amazingly, the MOFs can
accommodate a class of molecules and further distinguish among them with
Luminescent Properties and Applications of Metal-Organic Frameworks 69

characteristic different visible light emission for each different guest molecule.
They synthesized a 1,4,5,8-naphthalenediimide (NDI)-based interpenetrating MOF,
Zn2(1,4-BDC)2(dpNDI)·4DMF (1,4-BDC ¼ 1,4-benzenedicarboxylate; dpNDI,
N,N0 -di(4-pyridyl)-1,4,5,8-naphthalenediimide), which will undergo a dynamic
structural transformation to confine a class of aromatic volatile organic compounds
or tropospheric air pollutants. In the structure of the interpenetrating MOF
Zn2(1,4-BDC)2(dpNDI)·4DMF, one of the frameworks is intergrown at the central
position of the void that is created by the other framework. The void spaces can be
classified as two different environments: the characteristic slit between dpNDI and
1,4-BDC (site A) and the remaining void space (site B). The aromatic volatile
organic compounds (VOCs) such as benzene, toluene, xylene, anisole, and
iodobenzene can be introduced by immersion of a dry MOF into the liquid of
each aromatic VOC. The guest-exchanged MOF shows a significant framework
displacement, in which one framework slid along the main axis of one 1,4-BDC
molecule. The MOF Zn2(1,4-BDC)2(dpNDI) shows a very weak fluorescence
with a very low quantum yield and a short average lifetime. In contrast, all
VOC-incorporated Zn2(1,4-BDC)2(dpNDI) exhibit intense fluorescence in the
visible light region, and the emission color is dependent on the chemical substituent
of the aromatic guest. Furthermore, the normalized emission spectra exhibit a
gradual red shift as the electron-donating capability of VOCs increased upon
excitation at 370 nm (Fig. 10). The multicolor luminescence is attributed to the
enhanced naphthalenediimide-aromatic guest interaction through the induced-fit
structural transformation of the interpenetrating framework.

4.3.3 Gas and Vapor Sensor

Gas sensors have a wide range of application in the fields of aerodynamics,


environmental analysis, analytical chemistry, and biochemistry [214, 215]. Similar
to the sensing of small molecules and ions, the luminescent properties of MOFs can
also be perturbed by gases, thereby providing an effective means for sensing of gases.
Cyclometalated iridium complexes such as Ir(ppy)3 (ppy ¼ 2-phenylpyridine) are
highly efficient phosphorescent molecules and can be used to accurately detect
oxygen due to their MLCT phosphorescence and can be readily quenched by
molecules with a triplet ground state. Xie et al. recently synthesized a Zn–Ir
(2-phenylpyridine) MOF using Ir(ppy)3 derivatives [200], which possesses open
channels of 7.9  4.3 Å that are perpendicular to the (1, 1, 6) plane and contain
guest solvent molecules. To demonstrate the applicability for O2 sensing, the MOF
was gradually dosing in O2 from 0.05 to 1.0 atm, and the gradual decreases in
luminescence intensity at 538 nm were observed. Furthermore, the luminescence
measurements after alternating cycles of O2 dosing at 0.1 atm (30 s) and O2 removal
under vacuum (120 s) indicate that the luminescence is reversibly quenched by O2,
with <5% of the original luminescence lost after eight cycles.
Song et al. reported a lanthanide MOF, [Eu2L3(H2O)4]·3DMF (L ¼ 20 ,50 -bis
(methoxymethyl)-[1,10 :40 ,100 -terphenyl]-4,400 -dicarboxylate), for sensing of DMF
70 Y. Cui et al.

Fig. 10 Multicolor
luminescence of VOC
incorporated
Zn2(bdc)2(dpNDI). (a) The
resulting luminescence of
crystal powders of
Zn2(bdc)2(dpNDI)
suspended in each VOC
liquid after excitation at
365 nm using a commercial
ultraviolet lamp. (b) Height-
normalized luminescent
spectra of VOC
incorporated
Zn2(bdc)2(dpNDI) after
excitation at 370 nm.
Reprinted with the
permission from
[213]. Copyright 2011,
Macmillan Publishers Ltd

vapor [216]. The water-exchanged MOF displays a much weaker luminescence


under UV irradiation, while exhibits more than eightfold enhancement of lumines-
cence after incubation under DMF vapor (Fig. 11). This DMF-triggered turn-on
luminescence was rationalized by the DMF–ligand interactions that presumably
shift the excited state energy level of the ligand and thus facilitates the
ligand–lanthanide energy transfer process. The response rate of this MOF sensor
is quite fast, with 95% of turn-on and turn-off achieved within a few minutes and
10–20 s, respectively. This example offers the potential to build lanthanide-based
turn-on luminescent MOF sensors by proper ligand design, targeting analytes
through ligand–analyte interactions.
Luminescent Properties and Applications of Metal-Organic Frameworks 71

Fig. 11 Photograph (left) and emission spectra (right) of Eu2L3(H2O)4 (L ¼ 20 ,50 -bis
(methoxymethyl)-[1,10 :40 ,100 -terphenyl]-4,400 -dicarboxylate) before and after exposure to DMF
vapor. Reprinted with the permission from [216], Copyright 2013, John Wiley & Sons Ltd

Chemical sensors for rapid detection of explosives in the gas phase are attracting
increasing attention owing to homeland security, environment monitoring, and
humanitarian implications. Lan et al. reported a MOFs Zn2(bpdc)2(bpee)
(bpdc ¼ 4,40 -biphenyldicarboxylate; bpee ¼ 1,2-bipyridylethene) for detection of
both DNT and DMNB (2,3-dimethyl-2,3-dinitrobutane) with rapid response and
high sensitivity [157]. Within ten seconds, the fluorescence quench percentages of
the MOFs thin film reach almost the maxima for both DNT and DMNB (ca. 85%
and 84%, respectively). The sensitivity of the film for DMNB exceeds previously
reported materials based on conjugated polymer thin films. This was attributed to
the infinite 3D framework structure, the large surface area, and the pore confine-
ment of the analyte inside the molecular-sized cavities which facilitates stronger
interactions between the DMNB and the framework. Considering the very fast
responses and high sensitivity, it would be possible to produce a new and important
sensor for explosives using the MOFs.

4.3.4 Ratiometric Sensor

In most examples presented above, the sensing function is based on changes in the
intensity of one transition, which can be heavily affected by the quantity of the
luminophore, excitation power, and the drifts of the optoelectronic system. Thus,
the comparison of the emission of different samples based on the detected intensity
may lead to erroneous conclusions. Although the measurements of quantum yields
and/or lifetime are affected neither by the intensity of the excitation source nor by the
probe concentration, they require a relatively long time and the computational treat-
ment. The utilizing of the ratio between the intensity of two transitions of the same
luminescent material, instead of only one transition, can overcome the main
drawbacks of the intensity-based measurements of only one transition. The ratiometric
72 Y. Cui et al.

Fig. 12 (a) Optical microscopy image of ITQMOF-3-Eu under UV light. (b) Eu2 coordination
environment. (c) Eu1 coordination environment. C, gray; H, white–gray; N, blue; Eu, green; O,
red. (d) Intensity variation of the Eu2 5D0 ! 7F0 transition from pH value of 7.5 to 5; the inset
shows the linear variation of intensity ration of the 5D0!7F0 emissions with the pH value.
Reprinted with the permission from [138], Copyright 2009, John Wiley & Sons Ltd

luminescent sensor enables the sensing independent of the sample concentration and
of the drifts of the optoelectronic system such as lamp and detectors.
Harbuzaru et al. have given a detailed description of the use of the Eu3+-MOF
ITQMOF-3-Eu as ratiometric pH sensors. The MOF ITQMOF-3-Eu was
synthesized using 1,10-phenanthroline-2,9-dicarboxylic acid (H2PhenDCA) as
ligand, in which both carboxylate and phenanthroline moieties may coordinate to
the metal center [138]. The MOF ITQMOF-3-Eu can be described as a layered MOF
with two well-defined sheets, and each sheet contains one type of Eu3+ ions (Eu1 or
Eu2) in a different crystallographic position. Due to the two different Eu3+
environments, ITQMOF-3-Eu exhibits two nondegenerate 5D0 ! 7F0 emission
lines at 579.0 (Eu2) and 580.7 (Eu1) nm, respectively (Figure 12). The luminescence
features of each Eu3+ site may be rationalized in terms of the relationship between the
covalency of the Eu–(O,N) bonds and the nephelauxetic effect that influences the
energy of the 5D0 ! 7F0 transition: a site with more covalent Eu-(O,N) bonds has a
5
D0!7F0 transition at lower energy and a longer emission lifetime. Intriguingly,
only one of the two Eu3+ emitting sites is affected by the pH variation in the range
5–7.5; thus the sensor does not require external calibration. By following the
intensity ratio of the 5D0 ! 7F0 emissions of the two Eu3+ types, it is possible to
determine the pH of the solution. The high-emission quantum efficiency and
pH-sensing capability enable ITQMOF-3-Eu to be used in a new miniaturized pH
sensor prototype by combining the material with a commercial fiber optic 1.5 mm in
diameter.
Recently, Qian and Chen et al. reported the first ratiometric thermometer based on
luminescent MOF, Eu0.0069Tb0.9931-DMBDC (DMBDC ¼ 2,5-dimethoxy-1,4-
benzenedicarboxylate) [198]. The advantages of using luminescence over other
conventional thermometers are fast response, high sensitivity, noninvasive operation,
Luminescent Properties and Applications of Metal-Organic Frameworks 73

Fig. 13 (a) Emission spectra of Eu0.0069Tb0.9931-DMBDC recorded between 10 and 300 K;


(b) CIE chromaticity diagram showing the luminescence color of Eu0.0069Tb0.9931-DMBDC
at different temperatures. Reprinted with the permission from [198]. Copyright 2012, American
Chemical Society

and inertness to strong electric and magnetic fields. To date, most luminescence-based
thermometers rely on a single emission, whose accuracy can be heavily affected by
the quantity of the luminescent materials and excitation power. To overcome this
issue, the luminescent intensity ratio ITb/IEu of the mixed MOF instead of intensity
was utilized to realize the temperature sensing. Generally, the luminescent intensity of
Tb-DMBDC and Eu-DMBDC decreases gradually as the temperature increases due to
the thermal activation of nonradiative-decay pathways. However, the mixed Tb/Eu
MOF Eu0.0069Tb0.9931-DMBDC exhibits a significantly different temperature-
dependent luminescent behavior from those of Tb-DMBDC and Eu-DMBDC. The
emission intensity of the Tb3+ ions in Eu0.0069Tb0.9931-DMBDC decreases, while that
of the Eu3+ increases with the temperature (Fig. 13). The main emission bands of
613 (Eu3+) and 545 (Tb3+) nm were comparable at 10 K, while the emission of Eu3+
almost dominates the whole spectrum at 300 K. This is attributed to the phonon-
assisted Förster energy transfer, which results in an efficient energy transfer from Tb3+
to Eu3+ at high temperatures, as evidenced by luminescent lifetime measurements.
In addition, the tunable luminescence colors from the green yellow to red from 10 to
300 K allow directly visualizing the temperature change instantly and straight-
forwardly. Such a Tb/Eu MOF featuring temperature-dependent luminescence
enables itself to be an excellent candidate for self-referencing luminescent
thermometers as no further calibration of luminescence intensity is required.
The similar ratiometric temperature sensing was demonstrated by Hasegawa
et al. using a mixed MOF, [Tb0.99Eu0.01(hfa)3(dpbp)]n (hfa ¼ hexafluoro acetyl-
acetonato, dpbp ¼ 4,40 -bis(diphenylphosphoryl) biphenyl) in the range of 200 to
500 K [217]. The MOF exhibits the temperature-dependent emission. The emission
intensities at 543 nm of [Tb0.99Eu0.01(hfa)3(dpbp)]n decrease dramatically with
increasing temperature. In contrast, the emission intensities at 613 nm increase
74 Y. Cui et al.

Fig. 14 (a) Temperature-


dependent emission spectra
of mixed MOF
[Tb0.99Eu0.01(hfa)3(dpbp)]n
in the temperature range of
200–450 K; (b) Color
pictures of
[Tb0.99Eu0.01(hfa)3(dpbp)]n
under UV (365 nm)
irradiation; (c) Temperature
dependence of the energy
transfer efficiency (ηTb-Eu).
Reprinted with the
permission from [217],
Copyright 2013, John Wiley
& Sons Ltd

slightly (Fig. 14). This MOF exhibits brilliant green, yellow, orange, and red
photoluminescence under UV irradiation at 250, 300, 350, and 400 K, respectively.
Rocha and Carlos et al. also demonstrated the nanoparticles of MOF
Tb0.99Eu0.01(BDC)1.5(H2O)2 (BDC ¼ 1-4-benzendicarboxylate) as ratiometric lumi-
nescent nanothermometers in the physiological temperature (300–320 K) range
[218]. The nanoparticles were prepared by a reverse microemulsion method with
average length and diameter of 300 and 30 nm, respectively. Aqueous suspensions of
the nano-MOF display an emission quantum yield of 0.23 excitation at 320 nm and
relative sensitivity of 0.37% K1 at 318 K, suggesting the possibility of using nano-
MOFs to measure physiological temperatures. These MOF nanoparticles present a
considerable potential for applications of biological interest, particularly in multi-
modal such as thermal, optical, and magnetic resonance imaging.

4.4 Application in Biomedicine

Recently, the use of nanoscale lanthanide MOFs for biological and biomedical
applications has attracted increasing attentions. The interest of these functional
materials relies on the combination of the chemical or biofunctional behavior of
MOFs and the unique luminescence properties of lanthanide ions, such as high
photostability, long decay rates, large Stokes shifts, and narrow emission bands.
Besides their luminescent characteristics, lanthanide MOFs can possess paramag-
netic properties which help to increase the relaxation rate of water protons in the
Luminescent Properties and Applications of Metal-Organic Frameworks 75

Fig. 15 MR images of suspensions of Gd0.95(BDC)1.5(H2O)2 in water containing 0.1% xanthan gum


and luminescence images of ethanolic suspensions of Gd0.95(BDC)1.5(H2O)2, Gd0.95(BDC)1.5(H2O)2:
Eu0.05, and Gd0.95(BDC)1.5(H2O)2:Tb0.05. Reprinted with the permission from ref. [73]. Copyright
2006, American Chemical Society

tissues being imaged, making them useful as contrast agents in magnetic resonance
imaging (MRI) spectroscopy.
Multimodal imaging is a new imaging technique which combines more than one
imaging modality, such as X-ray, nuclear, ultrasound, computed tomography (CT),
MRI, and fluorescence imaging. Multimodal imaging is becoming more popular
because of its improved sensitivity, high resolution, and morphological visualization.
In particular, the combination of fluorescence imaging and MRI can ally the sensitivity
of the fluorescence component with the high degree of spatial resolution of MRI.
Lin et al. synthesized nanorods of MOF Gd(BDC)1.5(H2O)2 through a reverse
microemulsion method, which allows the control of the morphologies and sizes by
alteration of the water-surfactant ratio of the microemulsion system [73]. These
nanomaterials display large longitudinal relaxivities (R1) of 35.8 mM1 s1 and
transverse relaxivities (R2) of 55.6 mM1 s1 on a per Gd3+ basis and extraordinarily
large R1 of 1.6  107 mM1 s1 and R2 of 2.5  107 mM1 s1 on a per nanoparticle
basis. The level of R1 is unprecedented and at least an order of magnitude higher than
those of Gd3+-containing liposomes which have been shown to be effective target-
specific MRI contrast agents for cancer and cardiovascular disease. In addition, the
analogs Gd0.95(BDC)1.5(H2O)2:Eu0.05 and Gd0.95(BDC)1.5(H2O)2:Tb0.05 were also
synthesized. Ethanol suspensions of these materials are highly luminescent upon
UV excitation with characteristic red and green luminescence from Eu3+ and Tb3+,
respectively, suggesting that they can be used as potential contrast agents for multi-
modal imaging (Fig. 15).
The combination of molecular imaging and drug delivery in MOFs will lead to the
exciting possibility of MOF-based theranostics, in which imaging will be used to
guide, follow, and quantify the drug delivery process. Up to now, core–shell
architectures combining diverse functionalities and surface modifications into a
single nano-MOF have been designed as biosensing platforms for imaging,
targeting, diagnostics, and therapy. Lin et al. demonstrated a novel strategy of
delivering a fluorescence-imaging contrast agent and an anticancer drug by
postsynthetic modifications of a highly porous MOF [87]. They synthesized an
amino-functionalized iron–carboxylate MOF by incorporating 2-aminoterephthalic
76 Y. Cui et al.

Fig. 16 Functionalization of iron–carboxylate MOF with an optical imaging contrast agent


(BODIPY) and a prodrug of cisplatin (ESCP) through postsynthetic modifications. Reprinted
with the permission from [87]. Copyright 2009, American Chemical Society

acid (NH2–BDC). To load an optical imaging contrast agent, the MOF was
treated with (1,3,5,7-tetramethyl- 4,4-difluoro-8-bromomethyl-4-bora-3a,4a-diaza-s-
indacene (Br-BODIPY)) to yield a BODIPY-loaded particles with a loading capacity
of 5.6–11.6 wt %. In addition, the ethoxysuccinato-cisplatin (ESCP), which is a
prodrug of cisplatin, was also loaded into the MOF (Fig. 16). To improve the
biological stability, the BODIPY- and ESCP-loaded MOF particles were also coated
with silica shells using Na2SiO3 as the silica source to afford novel core–shell
nanostructures. Laser scanning confocal microscopy images show that the
BODIPY-loaded particles could cross the cell membrane and release the BODIPY
dye inside of the cell, thus the fluorescence was present in cells incubated with
BODIPY-loaded particles but absent in cells incubated without nanoparticles,
suggesting that the MOF is an efficient platform for delivering an optical contrast
agent. In addition, the modified BODIPY ligand was incubated with the cells, but no
fluorescence was observed at all within the cells, indicating that the core–shell
nanostructure was required for cell uptake. Treatment of HT-29 cells with
Luminescent Properties and Applications of Metal-Organic Frameworks 77

ESCP-loaded MOF particles shows appreciable cytotoxicity (IC50 ¼ 29 μM), which


is slightly less cytotoxic than cisplatin under the same conditions (IC50 ¼ 20 μM).
Further functionalization of silica-coated particles with silyl-derived c(RGDfK),
which is a cyclic peptide known to target the αvβ3 integrin, shows that these particles
has cytotoxicity (IC50 ¼ 21 μM) comparable to that of cisplatin. The generality of this
approach could be utilized for the design of a wide range of nanomedical devices for
theranostic applications.

5 Conclusions and Outlook

In the past decades, the growing research interest of chemists, physicists, and
materials scientists has greatly accelerated the development of luminescent
MOFs. Although luminescent MOFs are still in their infancy, the currently avail-
able results have unambiguously demonstrated that the design and construction of
MOFs for luminescent functionality is very active, and hundreds of papers
published in the last few years indicate that this interest is still increasing. The
basic principles and the promises of luminescent MOFs as multifunctional
materials have been established, as outlined in this chapter. On the one hand,
comprehensive experimental studies and screening on luminescence properties on
those reported and future-synthesized MOFs are still necessary to establish the
database of luminescent MOFs. On the other hand, some in-depth studies on the
mechanism of the luminescence behavior including the origin of the luminescence
and structure-luminescent property relationship need to be carried out through the
collaboration with theoretical scientists.
For construction of luminescent MOFs, it should be noted that the controllable
synthesis of the MOFs is also a key issue to be addressed because of various factors
affecting the formation process of the final products. The deliberate selection of
organic linkers is very important to ensure the construction of MOF materials with
luminescence. Rational control on structure, pore size, and functional sites within
luminescent MOFs will be enforced to target highly sensitive and selective
luminescent-sensing MOFs. Further efforts should be focused on the construction
of porous luminescent MOFs with multifunctional sites to collaboratively induce
their preferential and highly sensitive and selective binding with different ions. If
the luminescent sensing could be combined with the sieving function of the
micropores to allow small molecules to go through, while excluding larger species,
it would be possible to obtain some highly sensitive sensors for small molecules.
In most of the luminescent MOFs, the sensing functionality is fulfilled based on
the change of the luminescence intensity. The simple yet sophisticated ratiometric
sensing approach is one of the breakthroughs on the exploration of functional
luminescent MOF sensors. This approach makes the luminescent sensing indepen-
dent of the concentration of the sample and of the drifts of the optoelectronic system
including excitation source and detectors, thus overcoming the main drawbacks of
the intensity-based measurements of only one transition.
78 Y. Cui et al.

Nanoscale luminescent MOFs will expect to be fabricated into thin films for
their straightforward and instant sensing devices in the future. Additionally, nano-
scale luminescent MOFs have bright future in the fields of cell imaging, as well as
drug delivery monitoring and treatment, and will be extensively pursued for
theranostic nanomedicine in the near future. Clearly, it would be interesting to
utilize up-conversion luminescent MOFs in photodynamic therapy because the high
loading capacity and controllable drug release of MOFs can facilitate the delivery
and release of the photodynamic therapy drug. Moreover, the attractive multifunc-
tionality of MOFs makes it possible to combine the up-conversion luminescence
and drug delivery in a single component. In this regard, close multidisciplinary
collaboration among chemists, materials scientists, biomedical scientists, and
bioengineers will certainly facilitate the implementation of some promising lumi-
nescent MOF materials and technologies for practical applications for the
theranostic nanomedicine.

Acknowledgment This work was supported by the National Natural Science Foundation of
China (Nos. 51010002, 51272229, 51272231 and 51229201) and the Award AX-1730 from
Welch Foundation (BC).

References

1. Zhou H-C, Long JR, Yaghi OM (2012) Introduction to metal-organic frameworks. Chem Rev
112:673–674
2. Suh MP, Park HJ, Prasad TK et al (2012) Hydrogen storage in metal-organic frameworks.
Chem Rev 112:782–835
3. Sumida K, Rogow DL, Mason JA et al (2012) Carbon dioxide capture in metal-organic
frameworks. Chem Rev 112:724–781
4. Li J-R, Sculley J, Zhou H-C (2012) Metal-organic frameworks for separations. Chem Rev
112:869–932
5. Yoon M, Srirambalaji R, Kim K (2012) Homochiral metal-organic frameworks for asymmetric
heterogeneous catalysis. Chem Rev 112:1196–1231
6. Allendorf MD, Bauer CA, Bhakta RK et al (2009) Luminescent metal-organic frameworks.
Chem Soc Rev 38:1330–1352
7. Cui Y, Yue Y, Qian G et al (2012) Luminescent functional metal-organic frameworks.
Chem Rev 112:1126–1162
8. Chen BL, Xiang SC, Qian GD (2010) Metal-organic frameworks with functional pores for
recognition of small molecules. Acc Chem Res 43:1115–1124
9. Kurmoo M (2009) Magnetic metal-organic frameworks. Chem Soc Rev 38:1353–1379
10. Horcajada P, Gref R, Baati T et al (2012) Metal-organic frameworks in biomedicine.
Chem Rev 112:1232–1268
11. Murray LJ, Dinca M, Long JR (2009) Hydrogen storage in metal-organic frameworks.
Chem Soc Rev 38:1294–1314
12. Férey G (2008) Hybrid porous solids: past, present, future. Chem Soc Rev 37:191–214
13. Czaja AU, Trukhan N, Müller U (2009) Industrial applications of metal-organic frameworks.
Chem Soc Rev 38:1284
14. Zhang J-P, Zhang Y-B, Lin J-B et al (2012) Metal azolate frameworks: from crystal
engineering to functional materials. Chem Rev 112:1001–1033
Luminescent Properties and Applications of Metal-Organic Frameworks 79

15. Stock N, Biswas S (2012) Synthesis of metal-organic frameworks (MOFs): routes to various
mof topologies, morphologies, and composites. Chem Rev 112:933–969
16. Li H, Eddaoudi M, O’Keeffe M et al (1999) Design and synthesis of an exceptionally stable
and highly porous metal-organic framework. Nature 402:276–279
17. Chen B, Eddaoudi M, Hyde ST (2001) Interwoven metal-organic framework on a periodic
minimal surface with extra-large pores. Science 291:1021–1023
18. Rocha J, Carlos LD, Paz FAA et al (2011) Luminescent multifunctional lanthanides-based
metal-organic frameworks. Chem Soc Rev 40:926
19. Kreno LE, Leong K, Farha OK et al (2012) Metal-organic framework materials as chemical
sensors. Chem Rev 112:1105–1125
20. Jiang H-L, Xu Q (2011) Porous metal-organic frameworks as platforms for functional
applications. Chem Commun 47:3351–3370
21. Eddaoudi M, Kim J, Rosi N et al (2002) Systematic design of pore size and functionality in
isoreticular mofs and their application in methane storage. Science 295:469–472
22. Chen BL, Ma SQ, Zapata F et al (2007) Rationally designed micropores within a metal-
organic framework for selective sorption of gas molecules. Inorg Chem 46:1233–1236
23. Chen BL, Liang CD, Yang J et al (2006) A microporous metal-organic framework for
gas-chromatographic separation of alkanes. Angew Chem Int Ed 45:1390–1393
24. Wang RM, Zhang J, Li LJ (2010) A 2d metal-organic framework with a flexible cyclohexane-
1,2,5,6-tetracarboxylic acid ligand: synthesis, characterization and photoluminescent property.
J Mol Struct 970:14–18
25. Bai HY, Ma JF, Yang J et al (2010) Eight two-dimensional and three-dimensional metal
organic frameworks based on a flexible tetrakis(imidazole) ligand: synthesis, topological
structures, and photo luminescent properties. Cryst Growth Des 10:1946–1959
26. Ren P, Liu M-L, Zhang J et al (2008) 1d, 2d and 3d luminescent zinc(ii) coordination
polymers assembled from varying flexible thioether ligands. Dalton Trans 4711–4713
27. Chai X-C, Sun Y-Q, Lei R et al (2010) A series of lanthanide frameworks with a flexible
ligand, n, n0 -diacetic acid imidazolium, in different coordination modes. Cryst Growth Des
10:658–668
28. Li D-P, Zhou X-H, Liang X-Q et al (2010) Novel structural diversity of triazolate-based
coordination polymers generated solvothermally with anions. Cryst Growth Des 10:
2136–2145
29. Jing X, Meng H, Li G et al (2010) Construction of three metal-organic frameworks based on
multifunctional T-shaped tripodal ligands, H3pyimdc. Cryst Growth Des 10:3489–3495
30. Su Z, Chen S-S, Fan J et al (2010) Highly connected three-dimensional metal-organic
frameworks based on polynuclear secondary building units. Cryst Growth Des 10:3675–3684
31. Ma LQ, Abney C, Lin WB (2009) Enantioselective catalysis with homochiral metal-organic
frameworks. Chem Soc Rev 38:1248–1256
32. Li JR, Kuppler RJ, Zhou HC (2009) Selective gas adsorption and separation in metal-organic
frameworks. Chem Soc Rev 38:1477–1504
33. Long JR, Yaghi OM (2009) The pervasive chemistry of metal-organic frameworks.
Chem Soc Rev 38:1213–1214
34. Tian Y-Q, Chen Z-X, Weng L-H et al (2004) Two polymorphs of cobalt(II) imidazolate
polymers synthesized solvothermally by using one organic template n, n-dimethylacetamide.
Inorg Chem 43:4631–4635
35. Huang X-C, Lin Y-Y, Zhang J-P et al (2006) Ligand-directed strategy for zeolite-type metal-
organic frameworks: Zinc(II) imidazolates with unusual zeolitic topologies. Angew Chem Int
Ed 45:1557–1559
36. Fang Q, Zhu G, Xue M et al (2005) A metal-organic framework with the zeolite MTN
topology containing large cages of volume 2.5 nm3. Angew Chem Int Ed 44:3845–3848
37. Tian Y-Q, Cai C-X, Ji Y et al (2002) [Co5(im)10·2MB]1: a metal-organic open-framework
with zeolite-like topology Angew Chem Int Ed 41:1384–1386
80 Y. Cui et al.

38. Hayashi H, Côté AP, Furukawa H et al (2007) Zeolite a imidazolate frameworks. Nat Mater
6:501–506
39. Wang B, Côté AP, Furukawa H et al (2008) Colossal cages in zeolitic imidazolate
frameworks as selective carbon dioxide reservoirs. Nature 453:207–211
40. MacGillivray L (2010) Metal-organic frameworks design and application. Wiley, New York
41. Klimakow M, Klobes P, Thünemann AF et al (2010) Mechanochemical synthesis of metal-
organic frameworks: a fast and facile approach toward quantitative yields and high specific
surface areas. Chem Mater 22:5216–5221
42. Yuan W, Friščić T, Apperley D et al (2010) High reactivity of metal-organic frameworks
under grinding conditions: parallels with organic molecular materials. Angew Chem Int Ed
49:3916–3919
43. Braga D, Grepioni F, Maini L et al (2011) Solid-state reactivity of copper(I) iodide: lumines-
cent 2D-coordination polymers of CuI with saturated bidentate nitrogen bases. New J Chem
35:339–344
44. Friščić T, Reid DG, Halasz I et al (2009) Ion- and liquid-assisted grinding: Improved
mechanochemical synthesis of metal-organic frameworks reveals salt inclusion and anion
templating. Angew Chem Int Ed 49:712–715
45. Schlesinger M, Schulze S, Hietschold M et al (2010) Evaluation of synthetic methods for
microporous metal-organic frameworks exemplified by the competitive formation of
Cu2(BTC)3(H2O)3 and Cu2(BTC)(OH)(H2O). Microporous Mesoporous Mater 132:121–127
46. Yuan WB, Garay AL, Pichon A et al (2010) Study of the mechanochemical formation and
resulting properties of an archetypal mof: Cu3(BTC)(2) (btc ¼ 1,3,5-benzenetricarboxylate).
CrystEngComm 12:4063–4065
47. Pichon A, Lazuen-Garay A, James SL (2006) Solvent-free synthesis of a microporous metal-
organic framework. CrystEngComm 8:211–214
48. Kole GK, Koh LL, Lee SY et al (2010) A new ligand for metal-organic framework and
co-crystal synthesis: mechanochemical route to rctt-1,2,3,4-tetrakis-(40 -carboxyphenyl)-
cyclobutane. Chem Commun 46:3660–3662
49. Fujii K, Garay AL, Hill J et al (2010) Direct structure elucidation by powder X-ray diffraction
of a metal-organic framework material prepared by solvent-free grinding. Chem Commun
46:7572–7574
50. Son W-J, Kim J, Kim J, et al. (2008) Sonochemical synthesis of MOF-5. Chem Commun
44:6336–6338
51. Srivastava DN, Pol VG, Palchik O et al (2005) Preparation of stable porous nickel and cobalt
oxides using simple inorganic precursor, instead of alkoxides, by a sonochemical technique.
Ultrason Sonochem 12:205–212
52. Wang T, Lu XM, Han L et al (2008) Helical nanostructure of tubular metal-organic complex
synthesized by sonochemical process. Sci China Ser B Chem 51:971–975
53. Jung DW, Yang DA, Kim J et al (2010) Facile synthesis of MOF-177 by a sonochemical
method using 1-methyl-2-pyrrolidinone as a solvent. Dalton Trans 39:2883–2887
54. Sabouni R, Kazemian H, Rohani S (2010) A novel combined manufacturing technique for
rapid production of IRMOF-1 using ultrasound and microwave energies. Chem Eng J 165:
966–973
55. Jung D-W, Yang D-A, Kim J et al (2010) Facile synthesis of MOF-177 by a sonochemical
method using 1-methyl-2-pyrrolidinone as a solvent. Dalton Trans 39:2883
56. Prior TJ, Yotnoi B, Rujiwatra A (2011) Microwave synthesis and crystal structures of two
cobalt-4,40 -bipyridine-sulfate frameworks constructed from 1-D coordination polymers
linked by hydrogen bonding. Polyhedron 30:259–268
57. Khan NA, Haque E, Jhung SH (2010) Rapid syntheses of a metal-organic framework material
Cu3(BTC)2(H2O)3 under microwave: a quantitative analysis of accelerated syntheses.
PCCP 12:2625–2631
Luminescent Properties and Applications of Metal-Organic Frameworks 81

58. Haque E, Khan NA, Park JH et al (2010) Synthesis of a metal-organic framework material,
iron terephthalate, by ultrasound, microwave, and conventional electric heating: a kinetic
study. Chem Eur J 16:1046–1052
59. Lu CM, Liu J, Xiao KF et al (2010) Microwave enhanced synthesis of MOF-5 and its CO2
capture ability at moderate temperatures across multiple capture and release cycles.
Chem Eng J 156:465–470
60. Bux H, Liang FY, Li YS et al (2009) Zeolitic imidazolate framework membrane with molecular
sieving properties by microwave-assisted solvothermal synthesis. J Am Chem Soc 131:16000
61. Liu HK, Tsao TH, Zhang YT et al (2009) Microwave synthesis and single-crystal-to-single-
crystal transformation of magnesium coordination polymers exhibiting selective gas adsorp-
tion and luminescence properties. CrystEngComm 11:1462–1468
62. Yoo Y, Lai ZP, Jeong HK (2009) Fabrication of MOF-5 membranes using microwave-induced
rapid seeding and solvothermal secondary growth. Microporous Mesoporous Mater 123:
100–106
63. Sonnauer A, Stock N (2008) High-throughput and microwave investigation of rare earth
phosphonatoethanesulfonates-Ln(O3P-C2H4-SO3) (Ln ¼ Ho, Er, Tm, Yb, Lu, Y). J Solid
State Chem 181:3065–3070
64. Wang XF, Zhang YB, Huang H et al (2008) Microwave-assisted solvothermal synthesis of a
dynamic porous metal-carboxylate framework. Crystal Growth Design 8:4559–4563
65. Ni Z, Masel RI (2006) Rapid production of metal-organic frameworks via microwave-
assisted solvothermal synthesis. J Am Chem Soc 128:12394–12395
66. Centrone A, Harada T, Speakman S et al (2010) Facile synthesis of vanadium metal-organic
frameworks and their magnetic properties. Small 6:1598–1602
67. Carné A, Carbonell C, Imaz I et al (2011) Nanoscale metal-organic materials. Chem Soc Rev
40:291–305
68. Jahan M, Bao QL, Yang JX et al (2010) Structure-directing role of graphene in the synthesis
of metal-organic framework nanowire. J Am Chem Soc 132:14487–14495
69. Ma M, Zacher D, Zhang XN et al (2011) A method for the preparation of highly porous,
nanosized crystals of isoreticular metal-organic frameworks. Crystal Growth Design 11:
185–189
70. Ostermann R, Cravillon J, Weidmann C et al (2011) Metal-organic framework nanofibers via
electrospinning. Chem Commun 47:442–444
71. Li YS, Bux H, Feldhoff A et al (2010) Controllable synthesis of metal-organic frameworks:
from mof nanorods to oriented mof membranes. Adv Mater 22:3322
72. Lin W, Rieter WJ, Taylor KML (2009) Modular synthesis of functional nanoscale coordination
polymers. Angew Chem Int Ed 48:650–658
73. Rieter WJ, Taylor KML, An H et al (2006) Nanoscale metal-organic frameworks as potential
multimodal contrast enhancing agents. J Am Chem Soc 128:9024–9025
74. Taylor KML, Rieter WJ, Lin W (2008) Manganese-based nanoscale metal-organic
frameworks for magnetic resonance imaging. J Am Chem Soc 130:14358–14359
75. Xu H, Liu F, Cui Y et al (2011) A luminescent nanoscale metal-organic framework for
sensing of nitroaromatic explosives. Chem Commun 47:3153–3155
76. Taylor KML, Jin A, Lin W (2008) Surfactant-assisted synthesis of nanoscale gadolinium
metal-organic frameworks for potential multimodal imaging. Angew Chem Int Ed 47:
7722–7725
77. Zhou W, Wu H, Yildirim T (2008) Enhanced H2 adsorption in isostructural metal-organic
frameworks with open metal sites: strong dependence of the binding strength on metal ions.
J Am Chem Soc 130:15268–15269
78. Xiang S, Zhou W, Gallegos JM et al (2009) Exceptionally high acetylene uptake in a
microporous metal-organic framework with open metal sites. J Am Chem Soc 131:
12415–12419
82 Y. Cui et al.

79. Hasegawa S, Horike S, Matsuda R et al (2007) Three-dimensional porous coordination


polymer functionalized with amide groups based on tridentate ligand: selective sorption
and catalysis. J Am Chem Soc 129:2607–2614
80. Hwang YK, Hong D-Y, Chang J-S et al (2008) Amine grafting on coordinatively unsaturated
metal centers of MOFs: consequences for catalysis and metal encapsulation. Angew Chem Int
Ed 47:4144–4148
81. Xiang S, Zhou W, Zhang Z et al (2010) Open metal sites within isostructural metal-organic
frameworks for differential recognition of acetylene and extraordinarily high acetylene
storage capacity at room temperature. Angew Chem Int Ed 49:4615–4618
82. Cohen SM (2010) Modifying mofs: new chemistry, new materials. Chem Sci 1:32–36
83. Tanabe KK, Wang Z, Cohen SM (2008) Systematic functionalization of a metal-organic
framework via a postsynthetic modification approach. J Am Chem Soc 130:8508–8517
84. Wang Z, Cohen SM (2007) Postsynthetic covalent modification of a neutral metal-organic
framework. J Am Chem Soc 129:12368–12369
85. Nguyen JG, Cohen SM (2010) Moisture-resistant and superhydrophobic metal-organic
frameworks obtained via postsynthetic modification. J Am Chem Soc 132:4560–4561
86. Wang Z, Tanabe KK, Cohen SM (2010) Tuning hydrogen sorption properties of metal-
organic frameworks by postsynthetic covalent modification. Chem Eur J 16:212–217
87. Taylor-Pashow KML, Rocca JD, Xie Z et al (2009) Postsynthetic modifications of iron-
carboxylate nanoscale metal-organic frameworks for imaging and drug delivery. J Am Chem
Soc 131:14261–14263
88. Goto Y, Sato H, Shinkai S et al (2008) “Clickable” metal-organic framework. J Am Chem
Soc 130:14354–14355
89. Savonnet M, Bazer-Bachi D, Bats N et al (2010) Generic postfunctionalization route from
amino-derived metal-organic frameworks. J Am Chem Soc 132:4518–4519
90. Gadzikwa T, Farha OK, Malliakas CD et al (2009) Selective bifunctional modification of a
non-catenated metal-organic framework material via “click” chemistry. J Am Chem Soc
131:13613–13615
91. Berezin MY, Achilefu S (2010) Fluorescence lifetime measurements and biological imaging.
Chem Rev 110:2641–2684
92. Li X, Wang X-W, Zhang Y-H (2008) Blue photoluminescent 3d Zn(II) metal-organic
framework constructing from pyridine-2,4,6-tricarboxylate. Inorg Chem Commun 11:
832–834
93. Fang Q, Zhu G, Xue M et al (2006) Structure, luminescence, and adsorption properties of two
chiral microporous metal-organic frameworks. Inorg Chem 45:3582–3587
94. Shustova NB, McCarthy BD, Dinca M (2011) Turn-on fluorescence in tetraphenylethylene-
based metal-organic frameworks: an alternative to aggregation-induced emission. J Am
Chem Soc 133:20126–20129
95. Moore EG, Samuel APS, Raymond KN (2009) From antenna to assay: lessons learned in
lanthanide luminescence. Acc Chem Res 42:542–552
96. Bünzli J-CG, Piguet C (2002) Lanthanide-containing molecular and supramolecular
polymetallic functional assemblies. Chem Rev 102:1897–1928
97. Sabbatini N, Guardigli M, Lehn J-M (1993) Luminescent lanthanide complexes as photo-
chemical supramolecular devices. Coord Chem Rev 123:201–228
98. Binnemans K (2009) Lanthanide-based luminescent hybrid materials. Chem Rev 109:
4283–4374
99. Eliseeva SV, Pleshkov DN, Lyssenko KA et al (2010) Highly luminescent and triboluminescent
coordination polymers assembled from lanthanide β-diketonates and aromatic bidentate
o-donor ligands. Inorg Chem 49:9300–9311
100. de Lill DT, de Bettencourt-Dias A, Cahill CL (2007) Exploring lanthanide luminescence in
metal-organic frameworks: synthesis, structure, and guest-sensitized luminescence of a
mixed europium/terbium-adipate framework and a terbium-adipate framework. Inorg Chem
46:3960–3965
Luminescent Properties and Applications of Metal-Organic Frameworks 83

101. Soares-Santos PCR, Cunha-Silva L, Paz FAA et al (2008) Photoluminescent 3d lanthanide-


organic frameworks with 2,5-pyridinedicarboxylic and 1,4-phenylenediacetic acids.
Cryst Growth Des 8:2505–2516
102. Li C, Lin J (2010) Rare earth fluoride nano-/microcrystals: synthesis, surface modification
and application. J Mater Chem 20:6831
103. Carlos LD, Ferreira RAS, de Zea Bermudez V et al (2011) Progress on lanthanide-based
organic–inorganic hybrid phosphors. Chem Soc Rev 40:536
104. Yang J, Yue Q, Li G-D et al (2006) Structures, photoluminescence, up-conversion, and
magnetism of 2d and 3d rare-earth coordination polymers with multicarboxylate linkages.
Inorg Chem 45:2857–2865
105. Weng D, Zheng X, Jin L (2006) Assembly and upconversion properties of lanthanide
coordination polymers based on hexanuclear building blocks with (μ3-OH) bridges. Eur J
Inorg Chem 2006:4184–4190
106. Bünzli J-CG (2010) Lanthanide luminescence for biomedical analyses and imaging. Chem
Rev 110: 2729–2755
107. Mahata P, Ramya KV, Natarajan S (2008) Pillaring of CdCl2-like layers in lanthanide metal-
organic frameworks: synthesis, structure, and photophysical properties. Chem Eur J 14:
5839–5850
108. Xu J, Su W, Hong M (2011) A series of lanthanide secondary building units based metal-
organic frameworks constructed by organic pyridine-2,6-dicarboxylate and inorganic sulfate.
Cryst Growth Des 11:337–346
109. Chandler BD, Yu JO, Cramb DT et al (2007) Series of lanthanide-alkali metal-organic
frameworks exhibiting luminescence and permanent microporosity. Chem Mater 19:
4467–4473
110. Shi FN, Cunha-Silva L, Mafra L et al (2008) Interconvertable modular framework and
layered lanthanide(III)-etidronic acid coordination polymers. J Am Chem Soc 130:150–167
111. Zou J-P, Peng Q, Wen Z et al (2010) Two novel metal-organic frameworks (MOFs) with
(3,6)-connected net topologies: syntheses, crystal structures, third-order nonlinear optical and
luminescent properties. Cryst Growth Des 10:2613–2619
112. Wang G-H, Li Z-G, Jia H-Q et al (2009) Metal-organic frameworks based on the pyridine-
2,3-dicarboxylate and a flexible bispyridyl ligand: syntheses, structures, and photolumi-
nescence. CrystEngComm 11:292–297
113. Feng R, Jiang F-L, Chen L et al (2009) A luminescent homochiral 3d Cd(II) framework with a
threefold interpenetrating uniform net 86. Chem Commun 43:5296
114. Wang XW, Chen J-Z, Liu J-H (2007) Photoluminescent Zn(II) metalorganic frameworks
built from tetrazole ligand: 2d four-connected regular honeycomb (4363)-net. Cryst Growth
Des 7:1227–1229
115. An J, Shade CM, Chengelis-Czegan DA et al (2011) Zinc-adeninate metal-organic frame-
work for aqueous encapsulation and sensitization of near-infrared and visible emitting
lanthanide cations. J Am Chem Soc 133:1220–1223
116. Luo F, Batten SR (2010) Metal-organic framework (mof): Lanthanide(III)-doped approach
for luminescence modulation and luminescent sensing. Dalton Trans 39:4485
117. Park YK, Choi SB, Kim H et al (2007) Crystal structure and guest uptake of a mesoporous
metal-organic framework containing cages of 3.9 and 4.7 nm in diameter. Angew Chem Int
Ed 46:8230–8233
118. Buso D, Jasieniak J, Lay MDH et al (2012) Highly luminescent metal-organic frameworks
through quantum dot doping. Small 8:80–88
119. Fang Q-R, Zhu G-S, Jin Z et al (2007) Mesoporous metal-organic framework with rare etb
topology for hydrogen storage and dye assembly. Angew Chem Int Ed 46:6638–6642
120. Rodrı́guez-Diéguez A, Salinas-Castillo A, Sironi A et al (2010) A chiral diamondoid 3D
lanthanum metal-organic framework displaying blue-greenish long lifetime photolumi-
nescence emission. CrystEngComm 12:1876
84 Y. Cui et al.

121. Rao X, Huang Q, Yang X et al (2012) Color tunable and white light emitting Tb3+ and Eu3+
doped lanthanide metal-organic framework materials. J Mater Chem 22:3210–3214
122. Feng X, Wang L-Y, Zhao J-S et al (2010) Series of anion-directed lanthanide-rigid-flexible
frameworks: syntheses, structures, luminescence and magnetic properties. CrystEngComm
12:774
123. Zhang L-Z, Gu W, Li B et al (2007) {[Nd4(ox)4(NO3)2(OH)2(H2O)2]·5H2O}n: a porous 3d
lanthanide-based coordination polymer with a special luminescent property. Inorg Chem
46:622–624
124. Gl Z, Maury O, Thuéry P et al (2008) Structural diversity in neodymium bipyrimidine
compounds with near infrared luminescence: from mono- and binuclear complexes to
metal-organic frameworks. Inorg Chem 47:10398–10406
125. Zhu X, Lü J, Li X et al (2008) Syntheses, structures, near-infrared, and visible luminescence
of lanthanide-organic frameworks with flexible macrocyclic polyamine ligands.
Cryst Growth Des 8:1897–1901
126. Wang H-S, Zhao B, Zhai B et al (2007) Syntheses, structures, and photoluminescence of
one-dimensional lanthanide coordination polymers with 2,4,6-pyridinetricarboxylic acid.
Cryst Growth Des 7:1851–1857
127. Liao J-H, Tsai C-S, Lin T-K (2010) Syntheses, structural characterization and lumines-
cent properties of M2(ATPA)3(DMF)2(H2O)2 (M ¼ Nd, Sm, Eu, Gd, Tb, Dy; ATPA ¼
2-aminoterephthalate, DMF ¼ N, N-dimethylformamide). Inorg Chem Commun
13:286–289
128. Daiguebonne C, Kerbellec N, Guillou O et al (2008) Structural and luminescent properties of
micro- and nanosized particles of lanthanide terephthalate coordination polymers.
Inorg Chem 47:3700–3708
129. Han Y, Li X, Li L et al (2010) Structures and properties of porous coordination polymers
based on lanthanide carboxylate building units. Inorg Chem 49:10781–10787
130. de Lill DT, Gunning NS, Cahill CL (2005) Toward templated metalorganic frameworks:
synthesis, structures, thermal properties, and luminescence of three novel lanthanideadipate
frameworks. Inorg Chem 44:258–266
131. Chen B, Wang L, Xiao Y et al (2009) A luminescent metal-organic framework with Lewis
basic pyridyl sites for the sensing of metal ions. Angew Chem Int Ed 48:500–503
132. Xiao Y, Cui Y, Zheng Q et al (2010) A microporous luminescent metal-organic framework
for highly selective and sensitive sensing of Cu2+ in aqueous solution. Chem Commun 46:
5503
133. Chen B, Yang Y, Zapata F et al (2007) Luminescent open metal sites within a metal-organic
framework for sensing small molecules. Adv Mater 19:1693–1696
134. Ma D, Wang W, Li Y et al (2010) In situ 2,5-pyrazinedicarboxylate and oxalate ligands
synthesis leading to a microporous europium–organic framework capable of selective sensing
of small molecules. CrystEngComm 12:4372
135. Lin Z-J, Xu B, Liu T-F et al (2010) A series of lanthanide metal-organic frameworks based on
biphenyl-3,40 ,5-tricarboxylate: syntheses, structures, luminescence and magnetic properties.
Eur J Inorg Chem 2010:3842–3849
136. Zhu W-H, Wang Z-M, Gao S (2007) Two 3d porous lanthanidefumarateoxalate
frameworks exhibiting framework dynamics and luminescent change upon reversible de-
and rehydration. Inorg Chem 46:1337–1342
137. Harbuzaru BV, Corma A, Rey F et al (2008) Metal-organic nanoporous structures with
anisotropic photoluminescence and magnetic properties and their use as sensors.
Angew Chem Int Ed 47:1080–1083
138. Harbuzaru BV, Corma A, Rey F et al (2009) A miniaturized linear ph sensor based on a
highly photoluminescent self-assembled europium(III) metal-organic framework.
Angew Chem Int Ed 48:6476–6479
Luminescent Properties and Applications of Metal-Organic Frameworks 85

139. Yang X-P, Jones RA, Rivers JH et al (2007) Syntheses, structures and luminescent properties
of new lanthanide-based coordination polymers based on 1,4-benzenedicarboxylate (bdc).
Dalton Trans 35:3936
140. Sun Y-G, Yu W, Wang L et al (2010) Synthesis, structures, and luminescence of lanthanide
coordination polymers constructed from benzimidazole-5,6-dicarboxylate and oxalate
ligands. Inorg Chem Commun 13:479–483
141. Lu W-G, Jiang L, Feng X-L et al (2009) Three-dimensional lanthanide anionic
metalorganic frameworks with tunable luminescent properties induced by cation exchange.
Inorg Chem 48:6997–6999
142. Wong KL, Law GL, Yang YY et al (2006) A highly porous luminescent terbium–organic
framework for reversible anion sensing. Adv Mater 18:1051–1054
143. Chen B, Wang L, Zapata F et al (2008) A luminescent microporous metalorganic frame-
work for the recognition and sensing of anions. J Am Chem Soc 130:6718–6719
144. Xu H, Xiao Y, Rao X et al (2011) A metal-organic framework for selectively sensing of
PO43 anion in aqueous solution. J Alloys Compd 509:2552–2554
145. Lin Y-W, Jian B-R, Huang S-C et al (2010) Synthesis and characterization of three ytterbium
coordination polymers featuring various cationic species and a luminescence study of a
terbium analogue with open channels. Inorg Chem 49:2316–2324
146. Yang X, Rivers JH, McCarty WJ et al (2008) Synthesis and structures of luminescent ladder-
like lanthanide coordination polymers of 4-hydroxybenzenesulfonate. New J Chem 32:790
147. Guo X, Zhu G, Sun F et al (2006) Synthesis, structure, and luminescent properties of
microporous lanthanide metalorganic frameworks with inorganic rod-shaped building
units. Inorg Chem 45:2581–2587
148. Chen B, Yang Y, Zapata F et al (2006) Enhanced near-infrared-luminescence in an erbium
tetrafluoroterephthalate framework. Inorg Chem 45:8882–8886
149. Guo Z, Xu H, Su S et al (2011) A robust near infrared luminescent ytterbium metal-organic
framework for sensing of small molecules. Chem Commun 47:5551–5553
150. White KA, Chengelis DA, Zeller M et al (2009) Near-infrared emitting ytterbium metal-
organic frameworks with tunable excitation properties. Chem Commun 45:4506
151. Huang W, Wu D, Zhou P et al (2009) Luminescent and magnetic properties of lanthanide-
thiophene-2,5-dicarboxylate hybrid materials. Cryst Growth Des 9:1361–1369
152. Lee EY, Jang SY, Suh MP (2005) Multifunctionality and crystal dynamics of a highly stable,
porous metalorganic framework [Zn4O(NTB)2]. J Am Chem Soc 127:6374–6381
153. Feng PL, Perry Iv JJ, Nikodemski S et al (2010) Assessing the purity of metalorganic
frameworks using photoluminescence: MOF-5, ZnO quantum dots, and framework decom-
position. J Am Chem Soc 132:15487–15489
154. Zhang Z, Xiang S, Rao X et al (2010) A rod packing microporous metal-organic framework
with open metal sites for selective guest sorption and sensing of nitrobenzene.
Chem Commun 46:7205
155. Zhang Z, Xiang S, Zheng Q et al (2010) A rare uninodal 9-connected metal-organic
framework with permanent porosity. Cryst Growth Des 10:2372–2375
156. Hou L, Lin Y-Y, Chen X-M (2008) Porous metalorganic framework based on μ4-oxo
tetrazinc clusters: sorption and guest-dependent luminescent properties. Inorg Chem 47:
1346–1351
157. Lan A, Li K, Wu H et al (2009) A luminescent microporous metal-organic framework for the
fast and reversible detection of high explosives. Angew Chem Int Ed 121:2370–2374
158. Zhang C, Che Y, Zhang Z et al (2011) Fluorescent nanoscale zinc(II)-carboxylate coordination
polymers for explosive sensing. Chem Commun 47:2336–2338
159. Bauer CA, Timofeeva TV, Settersten TB et al (2007) Influence of connectivity and porosity
on ligand-based luminescence in zinc metal-organic frameworks. J Am Chem Soc 129:
7136–7144
160. Jiang H-L, Liu B, Xu Q (2010) Rational assembly of d10 metal-organic frameworks with
helical nanochannels based on flexible V-shaped ligand. Cryst Growth Des 10:806–811
86 Y. Cui et al.

161. Yang E-C, Zhao H-K, Ding B et al (2007) Four novel three-dimensional triazole-based zinc
(II) metal-organic frameworks controlled by the spacers of dicarboxylate ligands: Hydrothermal
synthesis, crystal structure, and luminescence properties. Cryst Growth Des 7:2009–2015
162. Guo H-D, Guo X-M, Batten SR et al (2009) Hydrothermal synthesis, structures, and lumi-
nescent properties of seven d10 metal-organic frameworks based on 9,9-dipropylfluorene-
2,7-dicarboxylic acid (H2DFDA). Cryst Growth Des 9:1394–1401
163. Chen F, Wu M-F, Liu G-N et al (2010) Zinc(II) and cadmium(II) coordination polymers
based on 3-(5H-tetrazolyl)benzoate ligand with different coordination modes: hydrothermal
syntheses, crystal structures and ligand-centered luminescence. Eur J Inorg Chem 2010:
4982–4991
164. Chen X-L, Gou L, Hu H-M et al (2008) New examples of metal coordination architectures of
4,40 -sulfonyldibenzoic acid: syntheses, crystal structure and luminescence. Eur J Inorg Chem
2008:239–250
165. Xue M, Zhu G, Zhang Y et al (2008) Rational design and control of the dimensions of channels
in a series of 3d pillared metal-organic frameworks: synthesis, structures, adsorption, and
luminescence properties. Cryst Growth Des 8:427–434
166. Jiang H-L, Tatsu Y, Lu Z-H et al (2010) Non-, micro-, and mesoporous metal-organic
framework isomers: reversible transformation, fluorescence sensing, and large molecule
separation. J Am Chem Soc 132:5586–5587
167. Qiu Y, Li Y, Peng G et al (2010) Cadmium metal-directed three-dimensional coordination
polymers: in situ tetrazole ligand synthesis, structures, and luminescent properties.
Cryst Growth Des 10:1332–1340
168. Liu H-J, Tao X-T, Yang J-X et al (2008) Three-dimensional metal-organic network architecture
with large π-conjugated indolocarbazole derivative: synthesis, supramolecular structure, and
highly enhanced fluorescence. Cryst Growth Des 8:259–264
169. Xue M, Zhu G, Li Y et al (2008) Structure, hydrogen storage, and luminescence properties of
three 3d metal-organic frameworks with NbO and PtS topologies. Cryst Growth Des 8:
2478–2483
170. Wang M-S, Guo S-P, Li Y et al (2009) A direct white-light-emitting metal-organic frame-
work with tunable yellow-to-white photoluminescence by variation of excitation light. J Am
Chem Soc 131:13572–13573
171. Song L, Du S-W, Lin J-D et al (2007) A 3d metal-organic framework with rare 3-fold
interpenetrating dia-g nets based on silver(I) and novel tetradentate imidazolate ligand:
synthesis, structure, and possible ferroelectric property. Cryst Growth Des 7:2268–2271
172. Wu H-C, Thanasekaran P, Tsai C-H et al (2006) Self-assembly, reorganization, and
photophysical properties of silver(I)-schiff-base molecular rectangle and polymeric array
species. Inorg Chem 45:295–303
173. Bai Y, Gao H, Dang D-B et al (2010) A series of metal-organic frameworks based on
polydentate schiff-base ligands derived from benzil dihydrazone: synthesis, crystal structures
and luminescent properties. CrystEngComm 12:1422
174. Zhang T, Ji C, Wang K et al (2010) First halogen anion-bridged (MMX)n-type
one-dimensional coordination polymer built upon d10d10 dimers. Inorg Chem 49:
11069–11076
175. Jiang H, Ma J-F, Zhang W-L et al (2008) Metal-organic frameworks containing flexible bis
(benzimidazole) ligands. Eur J Inorg Chem 2008:745–755
176. Bai Y, He G-J, Zhao Y-G et al (2006) Porous material for absorption and luminescent
detection of aromatic molecules in water. Chem Commun 42:1530
177. Li M-X, Wang H, Liang S-W et al (2009) Solvothermal synthesis and diverse coordinate
structures of a series of luminescent copper(I) thiocyanate coordination polymers based on
n-heterocyclic ligands. Cryst Growth Des 9:4626–4633
178. Liu X, Huang K-L (2009) A 12-connected dodecanuclear copper cluster with yellow lumi-
nescence. Inorg Chem 48:8653–8655
Luminescent Properties and Applications of Metal-Organic Frameworks 87

179. Xia J, Zhang Z-J, Shi W et al (2010) Copper(I) cyanide coordination polymers constructed
from bis(pyrazole-1-yl)alkane ligands: observation of the odd-even dependence in the
structures. Cryst Growth Des 10:2323–2330
180. He J, Yin Y-G, Wu T et al (2006) Design and solvothermal synthesis of luminescent copper
(I)-pyrazolate coordination oligomer and polymer frameworks. Chem Commun 42:2845
181. Xiong S, Wang S, Tang X et al (2011) Four new metal-organic frameworks constructed
from H2DBTDC-O2 (H2DBTDC-O2¼dibenzothiophene-5,50 -dioxide-3,7-dicarboxylic acid)
ligand with guest-responsive photoluminescence. CrystEngComm 13:1646–1653
182. Wei Y, Yu Y, Wu K (2008) Highly stable five-coordinated Mn(II) polymer [Mn(hbidc)]n
(hbidc ¼ 1H-benzimidazole-5,6-dicarboxylate): crystal structure, antiferromegnetic prop-
erty, and strong long-lived luminescence. Cryst Growth Des 8:2087–2089
183. Bo Q-B, Sun G-X, Geng D-L (2010) Novel three-dimensional pillared-layer Ln(III)Cu
(I) coordination polymers featuring spindle-shaped heterometallic building units. Inorg Chem
49:561–571
184. Sun Y-Q, Zhang J, Yang G-Y (2006) A series of luminescent lanthanide–cadmium–organic
frameworks with helical channels and tubes. Chem Commun 45:4700
185. Zhao X-Q, Zhao B, Shi W et al (2009) Structures and luminescent properties of a series of
ln–ag heterometallic coordination polymers. CrystEngComm 11:1261
186. Liu W, Li Z, Wang N et al (2011) A new family of 3d heterometallic 3d–4f organodisulfonate
complexes based on the linkages of 2D [ln(nds)(H2O)]+ layers and [Cu(ina)2] chains.
CrystEngComm 13:138
187. Chandler BD, Cramb DT, Shimizu GKH (2006) Microporous metalorganic frameworks
formed in a stepwise manner from luminescent building blocks. J Am Chem Soc 128:
10403–10412
188. Zhao B, Chen X-Y, Cheng P et al (2004) Coordination polymers containing 1d channels as
selective luminescent probes. J Am Chem Soc 126:15394–15395
189. Zhao B, Chen X-Y, Chen Z et al (2009) A porous 3d heterometal-organic framework
containing both lanthanide and high-spin Fe(II) ions. Chem Commun 45:3113
190. Gu X, Xue D (2007) Self-assembly of 3-d 4d-4f coordination frameworks based on 1-d
inorganic heterometallic chains and linear organic linkers. CrystEngComm 9:471
191. Peng G, Qiu Y-C, Liu Z-H et al (2008) Construction of three isostructural 3d–4f microporous
coordination frameworks based on mixed nicotinate and oxalate ligands. Inorg Chem
Commun 11:1409–1411
192. Wang P, Ma J-P, Dong Y-B et al (2007) Tunable luminescent lanthanide coordination
polymers based on reversible solid-state ion-exchange monitored by ion-dependent photo-
induced emission spectra. J Am Chem Soc 129:10620–10621
193. Jin J, Niu S, Han Q et al (2010) Synthesis and structure of a series of new luminescent Ag–Ln
coordination polymers and the influence of the introduction of an Ag(I) ion on nir lumines-
cence from the Ln(III) centre. New J Chem 34:1176
194. Zhao B, Zhao XQ, Chen Z et al (2008) Structures and near-infrared luminescence of unique
4d–4f heterometal-organic frameworks (hmof). CrystEngComm 10:1144
195. Prasad TK, Rajasekharan MV (2009) Cerium(IV)-lanthanide(III)-pyridine-2,6-dicarboxylic
acid system: coordination salts, chains, and rings. Inorg Chem 48:11543–11550
196. Guo H, Zhu Y, Qiu S et al (2010) Coordination modulation induced synthesis of nanoscale
eu1-xtbx-metal-organic frameworks for luminescent thin films. Adv Mater 22:4190–4192
197. Liu K, You H, Zheng Y et al (2010) Facile and rapid fabrication of metal-organic framework
nanobelts and color-tunable photoluminescence properties. J Mater Chem 20:3272
198. Cui Y, Xu H, Yue Y et al (2012) A luminescent mixed-lanthanide metal-organic framework
thermometer. J Am Chem Soc 134:3979–3982
199. White KA, Chengelis DA, Gogick KA et al (2009) Near-infrared luminescent lanthanide mof
barcodes. J Am Chem Soc 131:18069–18071
200. Xie Z, Ma L, deKrafft KE et al (2010) Porous phosphorescent coordination polymers for
oxygen sensing. J Am Chem Soc 132:922–923
88 Y. Cui et al.

201. Stylianou KC, Heck R, Chong SY et al (2010) A guest-responsive fluorescent 3d microporous


metalorganic framework derived from a long-lifetime pyrene core. J Am Chem Soc 132:
4119–4130
202. Sava DF, Rohwer LES, Rodriguez MA et al (2012) Intrinsic broad-band white-light emission
by a tuned, corrugated metal-organic framework. J Am Chem Soc 134:3983–3986
203. Wibowo AC, Vaughn SA, Smith MD et al (2010) Novel bismuth and lead coordination
polymers synthesized with pyridine-2,5-dicarboxylates: two single component “white” light
emitting phosphors. Inorg Chem 49:11001–11008
204. Yang E-C, Li J, Ding B et al (2008) An eight-connected 3d lead(ii) metal-organic framework
with octanuclear lead(ii) as a secondary building unit: synthesis, characterization and lumi-
nescent property. CrystEngComm 10:158–161
205. Hu R, Cai H, Luo J (2011) Synthesis, structure and luminescent property of a pillared-layer
coordination polymer [pb(bpdc)] (bpdc ¼ 4,40 -biphenyldicarboxylate). Inorg Chem Commun
14:433–436
206. D’Andrade BW, Forrest SR (2004) White organic light-emitting devices for solid-state
lighting. Adv Mater 16:1585
207. Eliseeva SV, Bünzli J-CG (2010) Lanthanide luminescence for functional materials and
bio-sciences. Chem Soc Rev 39:189
208. Dang S, Zhang J-H, Sun Z-M (2012) Tunable emission based on lanthanide(III) metal-
organic frameworks: an alternative approach to white light. J Mater Chem 22:8868
209. Kozlowski H, Janicka-Klos A, Brasun J et al (2009) Copper, iron, and zinc ions homeostasis
and their role in neurodegenerative disorders (metal uptake, transport, distribution and
regulation). Coord Chem Rev 253:2665–2685
210. Que EL, Domaille DW, Chang CJ (2008) Metals in neurobiology: probing their chemistry
and biology with molecular imaging. Chem Rev 108:1517–1549
211. Qiu Y, Deng H, Mou J et al (2009) In situ tetrazole ligand synthesis leading to a microporous
cadmium–organic framework for selective ion sensing. Chem Commun 45:5415
212. Xu H, Rao X, Gao J et al (2012) A luminescent nanoscale metal-organic framework with
controllable morphologies for spore detection. Chem Commun 48:7377–7379
213. Takashima Y, Martı́nez VM, Furukawa S et al (2011) Molecular decoding using lumines-
cence from an entangled porous framework. Nat Comms 2:168
214. Katz MJ, Ramnial T, Yu H-Z et al (2008) Polymorphism of Zn[Au(CN)2]2and its luminescent
sensory response to NH3 vapor. J Am Chem Soc 130:10662–10673
215. Habibagahi A, Mébarki Y, Sultan Y et al (2009) Water-based oxygen-sensor films. ACS Appl
Mater Interfaces 1:1785–1792
216. Li Y, Zhang S, Song D (2013) A luminescent metal-organic framework as a turn-on sensor
for DMF vapor. Angew Chem Int Ed 52:710–713
217. Miyata K, Konno Y, Nakanishi T et al (2013) Chameleon luminophore for sensing
temperatures: control of metal-to-metal and energy back transfer in lanthanide coordination
polymers. Angew Chem Int Ed 52:6413–6416
218. Cadiau A, Brites CD, Costa PM et al (2013) Ratiometric nanothermometer based on an
emissive Ln-organic framework. ACS Nano 7:7213–7218.
Struct Bond (2014) 157: 89–104
DOI: 10.1007/430_2013_131
# Springer-Verlag Berlin Heidelberg 2013
Published online: 7 September 2013

Metal-Organic Frameworks for


Photocatalysis

Teng Zhang and Wenbin Lin

Abstract Metal-organic frameworks (MOFs) have recently been investigated as a


platform for developing single-site photocatalysts and multifunctional photo-
catalytic assemblies. This chapter aims to summarize the recent progress on
photocatalysis with MOFs, including hydrogen evolution, carbon dioxide reduction,
and degradation and transformation of organic compounds.

Keywords CO2 reduction  Metal-organic frameworks  Photocatalysis  Water


splitting

Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
2 Hydrogen Evolution with MOFs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
3 Carbon Dioxide Reduction with MOFs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
4 Photocatalytic Degradation of Organic Compounds with MOFs . . . . . . . . . . . . . . . . . . . . . . . . . . 95
5 Photocatalytic Organic Transformations with MOFs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
6 Conclusions and Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

T. Zhang and W. Lin (*)


Department of Chemistry, The University of Chicago, 929 E. 57th Street, Chicago, IL 60637,
USA
e-mail: wenbinlin@uchicago.edu
90 T. Zhang and W. Lin

1 Introduction

With the dwindling fossil fuel reserve and increasing global energy demand, solar
energy is the best alternative energy source. Significant efforts have been devoted to
harvesting and storing energy from the sunlight. Photochemical processes with net
energy uphill reactions (ΔG > 0), such as water splitting and carbon dioxide
reduction, have been proposed to transform and store sunlight energy in the form
of chemical bonds. Green plants have developed effective but complex systems to
practice these transformations, known as photosynthesis, to convert carbon dioxide
and water into carbohydrates. Inspired by natural photosynthesis, scientists have
strived to develop inorganic materials and devices for these photochemical
transformations. Since the first report of UV light-driven water splitting in
1972 [1], different materials and devices have been demonstrated [2–5], yet it
still remains a great challenge to develop a stable, effective, and affordable system
that can utilize the full solar spectrum for solar energy conversion.
Chemists have also tried to mimic the molecular photochemical systems in green
plants. Various molecular dyes as well as catalysts have been developed
[6–11]. There are however no efficient strategies to combine these functional
components effectively into hierarchical assemblies for artificial photosynthesis.
Metal-organic frameworks (MOFs) provide a potential solution to this challenge.
Constructed from well-defined metal/metal cluster nodes and molecular building
blocks, MOFs have shown potential applications in many different areas [12–19]
and offer a possibility to incorporate different functional components into a single
solid. In particular, MOFs allow hierarchical assembly of multiple functionalities
into a single solid material to potentially achieve artificial photosynthesis.
Light also serves as a means to activate organic molecules to undergo
transformations that cannot be carried out under typical thermally activated reaction
conditions. Although photocatalysis has long been proposed as an important method-
ology, the application is limited because most organic substrates only absorb UV
photons, leaving the majority of the solar spectrum unused. Molecular dyes can be
used to absorb lower-energy photons to activate organic substrates, leading to many
novel photocatalytic reactions driven with visible light [20–24]. Photoactive MOFs
can also serve as single-site photocatalysts and take advantage of the solid nature of
MOF catalysts to facilitate recovery and reuse of expensive photocatalysts thereby
reducing contamination by the photocatalysts which usually contain heavy metals. In
this chapter, we will summarize recent progress on photocatalysis with MOFs,
including water splitting, carbon dioxide reduction, and organic transformations.

2 Hydrogen Evolution with MOFs

During the past few decades, semiconducting materials and molecular complexes
have been developed as photosensitizers or photocatalysts for splitting water into
hydrogen and oxygen. Several MOFs have also been reported to serve as
Metal-Organic Frameworks for Photocatalysis 91

Fig. 1 Proposed photocatalytic cycle of hydrogen evolution from water using the [Ru2( p-BDC)2]n,
Ru(bpy)32+, MV2+, and Na2EDTA combination under visible light irradiation. Reprinted with
permission from [25]. Copyright 2009 Royal Society of Chemistry

photosensitizers or catalysts for sacrificial proton reduction reactions in the past few
years. In 2009, Mori and coworkers reported a porous MOF [Ru2( p-BDC)2]n
(1, p-BDC ¼ p-benzenedicarboxylate) that catalyzes visible light-driven proton
reduction with the Ru(bpy)32+/MV2+/EDTA system [25]. 1 adopts in a 2-D square
grid structure constructed from Ru2(COO)4 paddle-wheel building blocks and
linear p-BDC linkers, with 1-D channels running perpendicular to the 2-D sheets.
In the system, 1 serves as the catalyst, Ru(bpy)32+ acts as a photosensitizer, MV2+
acts as an electron relay, and Na2EDTA acts as a sacrificial reductant. The catalytic
proton reduction reaction gives a turnover number (TON) of 8.16 per Ru atom in a
4 h period under irradiation. The proposed catalytic cycle starts with the electron
transfer between the excited state of Ru(bpy)32+ and the electron relay MV2+ to
yield Ru(bpy)33+ and MV+, while the former is reduced back to Ru(bpy)32+ by
Na2EDTA and the latter transfers an electron to the catalyst 1 to reduce proton to
hydrogen (Fig. 1). Although the proposed mechanism seems plausible, the authors
provided little experimental evidence to support this proposal.
The Zr-based MOFs UiO-66 and UiO-66(NH2), first developed by Lillerud and
coworkers [26], have also been reported for photocatalytic water reduction. In
2010, Garcı́a and coworkers reported hydrogen evolution catalyzed by the MOFs
with a UV irradiation (>300 nm) and methanol as the sacrificial electron donor
[27]. Platinum nanoparticles were also added as a cocatalyst. The presence of amino
group shifts the absorption band to >300 nm and makes the photocatalysis possible.
However, the need for UV photons presents a significant drawback for this experi-
mental design.
In early 2012, Lin and coworkers reported a MOF assembly containing molecular
phosphor building blocks and platinum nanoparticles for hydrogen evolution
[28]. Iridium-based molecular photosensitizers bis(4-phenyl-2-pyridine)(5,50 -
dicarboxylate)-2,20 -bipyridine)-iridium(III) chloride (L1) or bis(4-phenyl-2-pyridine)
(5,50 -di(4-phenylcarboxylate)-2,20 -bipyridine)-iridium(III) chloride (L2) were used as
building blocks, in combination with Zr6(O)4(OH)4 SBUs, to construct water-stable
MOFs 2 or 3 (Fig. 2). Platinum nanoparticles were deposited in the MOF channels via
in situ MOF-mediated photoreduction of platinum precursor (Fig. 3). Higher nano-
particle loading is found in 3 than 2 as 3 has a larger void space and a higher phosphor
concentration. Both Pt@MOF assemblies exhibited high catalytic activity for proton
reduction with triethylamine as the sacrificial reductant. Pt@2 and Pt@3 exhibited a
92 T. Zhang and W. Lin

Fig. 2 Synthesis of phosphorescent Zr-carboxylate MOFs and subsequent loading of Pt NPs


inside MOFs 2 and 3. Reprinted with permission from [28]. Copyright 2012 American Chemical
Society

Fig. 3 (a) TEM images of Pt@3. (b) The synergistic photocatalytic hydrogen generation process
via photoinjection of electron from the light-harvesting MOF on to the Pt NPs. (c) HRTEM images
of a powdery sample of Pt@4 that shows the lattice fringes of the Pt particles, with d-spacing
matching that of the Pt{111} plane. Reprinted with permission from [28]. Copyright 2012
American Chemical Society
Metal-Organic Frameworks for Photocatalysis 93

TON of 3,400 and 7,000 per Ir for hydrogen evolution over an illumination time of
48 h, respectively. These TONs are almost five times as high as the homogeneous
control. It is believed that facile electron transfer between the photoreduced
Ir-complexes and the Pt nanoparticles enhances the catalytic performance of the
Pt@MOFs over their homogeneous counterparts. The catalysts can be recycled and
reused for at least three times though significant leaching of the iridium complexes
was also observed after the 48 h reactions.
Porphyrin-based ligands were also used to construct photoactive MOFs due to
their interesting photophysical properties. In 2012, Rosseinsky and coworkers
reported a water-stable porphyrin MOF (Al-PMOF) and its catalytic activity for
proton reduction [29]. They first tested the Al-PMOF/MV2+/EDTA/Pt system,
which is analogous to the one used by Mori et al. [25] but using Al-PMOF instead
of Ru(bpy)32+ as the photosensitizer. However, very low activity was observed.
When MV2+ was removed from the system, the hydrogen generation activity
increased by almost one order of magnitude and an approximate TON of 0.7
H2/porphyrin was achieved over a period of 6 h. The authors believed that the
slow diffusion of methyl viologen through the MOF channels results in ineffective
electron transfer and leads to low activity. Stability of the Al-PMOF was supported
by powder XRD after reaction and further confirmed by the absence of porphyrin
leaching into the solution after reaction.

3 Carbon Dioxide Reduction with MOFs

Photochemical reduction of CO2 is another important strategy for solar energy


conversion. Not only does it harvest energy from sunlight to produce valuable
organic compounds but also helps to reduce the CO2 level in the atmosphere. In
natural photosynthesis, CO2 is reduced by photochemically generated NADPH in
photosystem I [30]. Scientists have developed a number of molecular compounds,
semiconductors, and metal-incorporated zeolites for artificial photochemical CO2
reduction [31–34].
Lin and coworkers reported in 2011 the synthesis of MOF 4 by doping
[ReI(CO)3(5,50 -dcbpy)Cl] (L3), a derivative of the CO2 reduction catalyst
ReI(CO)3(bpy)X [31, 36, 37], into the UiO-67 framework at a loading of 4 wt%
[35]. The synthesis of 4 took advantage of the matching ligand lengths of
[ReI(CO)3(5,50 -dcbpy)Cl] and bpdc, while 4 showed the same PXRD pattern as
UiO-67 (Scheme 1 and Fig. 4c). Photocatalytic CO2 reduction was carried out in a
CO2-saturated acetonitrile solution with triethylamine as the sacrificial reductant and
4 as the catalyst. 4 selectively reduced CO2 to CO with a TON of 10.9 over a period of
12 h, almost three times higher than the homogeneous complex (Fig. 4a). The higher
activity was believed to be a result of blocking the bimolecular decomposition due to
site isolation of the [ReI(CO)3(5,50 -dcbpy)Cl] molecules. Moreover, 4 exhibited high
selectivity towards CO. The molar amount of CO was ten times as much as H2
generated in the first 6 h, and no formic acid or methanol was detected during the
94 T. Zhang and W. Lin

CO2H CO2H

CO
N CO DMF
ZrCl4 + + Re Zr6(µ3-O)4(µ3-OH)4(bpdc)6-x(L3)x
N CO 100oC
5
Cl
CO2H CO2H
H2bpdc H2L3

Scheme 1 Synthesis of 4 by doping [ReI(CO)3(5,50 ,-dcbpy)Cl] into the UiO-67 framework

Fig. 4 (a) Plots of CO evolution turnover number versus time in the photocatalytic CO2 reduction
with 4 and the homogeneous control. (b) FT-IR of 4 before and after photocatalysis. (c) PXRD
patterns of 4 before and after photocatalysis. The simulated PXRD for the UiO-67 is also shown.
(d) UV–vis diffuse reflectance spectra of 4 before and after photocatalysis. Reprinted with
permission from [35]. Copyright 2011 American Chemical Society

reaction. The recovered solid was found to be crystalline as shown by PXRD (Fig. 4c)
but became inactive for CO generation. The loss of MLCT absorption band (Fig. 4d)
and CO stretching vibrations (Fig. 4b) indicate that the detachment of Re-carbonyl
moieties from the MOF backbones is responsible for the loss of catalytic activity. In
addition, since the bimolecular pathway for CO2 reduction is prohibited in 4 due to
site isolation, the observed catalytic activity provides strong support for the
Metal-Organic Frameworks for Photocatalysis 95

Fig. 5 Absorption spectra and proposed mechanism for photocatalytic CO2 reduction of
MIL-125(Ti) and NH2-MIL-125(Ti). Reprinted with permission from [38]. Copyright 2012
Wiley-VCH

unimolecular pathway of photochemical CO2 reduction. This work thus not only
demonstrated the possibility of using MOFs for photocatalytic CO2 reduction but also
provided mechanistic insights for this reaction.
Li and coworkers reported in 2012 a photoactive MOF Ti8O8(OH)4(bdc-NH2)6
(NH2-MIL-125(Ti)) as the CO2 reduction catalyst under visible light [38]. The
MOF shows a visible absorption band that extends to 550 nm as a result of
introducing the amino group (Fig. 5). In the photocatalytic reaction, CO2 was
reduced into HCO2- photochemically in acetonitrile with triethanolamine (TEOA)
as the sacrificial reducing agent. It was proposed that, upon illumination, a long-
lived charge transfer excited state was generated and reduced by TEOA to give a
Ti3+ center which in turn reduced CO2 to give HCO2-. NH2-MIL-125(Ti) showed
modest photocatalytic activity for photochemical CO2 reduction, with a TON of
0.03 per Ti achieved in 10 h.

4 Photocatalytic Degradation of Organic Compounds with


MOFs

Photocatalytic degradation of organic compounds is the most explored area in MOF


photocatalysis [39–51]. MOF-5, with the framework composition of Zn4O(BDC)6
[52], was reported as a photocatalyst by Garcia et al. [39, 40]. MOF-5 has an
absorption spectrum that extends to 400 nm. It showed comparable activity for
oxidative degradation of phenol as that of the conventional P-25 TiO2 standard
[40]. Interestingly, Garcia and coworkers found that MOF-5 exhibits reverse shape
selectivity when using different phenol substrates [39]. The more bulky phenol
substrate, 2,6-di-t-butylphenol, exhibits a higher degradation rate compared to
phenol. The reverse selectivity was explained based on the argument that unlike
96 T. Zhang and W. Lin

Fig. 6 UV-vis absorption spectra (a) and photograph (b) of methyl orange solution degraded by
UTSA-38 after the UV–vis light irradiation for different times. Reprinted with permission from
[45]. Copyright 2011 Royal Society of Chemistry

smaller molecules which can diffuse into the MOF freely, large molecules remain
on the external surface of the MOF where the rate of degradation is faster.
Photoactive MOFs have also been investigated for degradation of organic dyes
such as methylene blue (MB) or rhodamine B (RhB). Ma et al. reported in 2012 a
MOF constructed from two different ligands tetrakis[4-(carboxyphenyl)-
oxamethyl]methane acid (H4L4) and tetrakis(imidazol-1-ylmethyl)methane (L5)
with the formula [Co2(L4)(L5)]·4.25H2O (5) [49]. 5 has an absorption edge of
3.78 eV and exhibits catalytic activity for degradation of the dyes MB, RhB, and
X3B. Under UV irradiation in the presence of 5, MB degraded to 5.25% after
75 min, while RhB degraded to 34% and X3B degraded to 43% after 10 h,
respectively. Moreover, 5 also catalyzed the degradation of MB under visible
irradiation. After 5 h of visible irradiation, 49.6% of MB was degraded in the
presence of 5.
Chen et al. reported in 2011 the synthesis of a twofold interpenetrated MOF
Zn4O(2,6-naphthalenedicarboxylate)3(DMF)1.5(H2O)0.5•4DMF•7.5H2O (UTSA-38)
and its catalytic activity for degrading methyl orange (MO) [45]. UTSA-38 has a
band gap of 2.38 eV and exhibits catalytic activity under UV–visible light (Fig. 6).
Li et al. reported in 2012 a family of MOFs constructed from the ligands
1,2,4,5-benzenetetracarboxylic acid (btec) and 4,40 -bis(1-imidazolyl)biphenyl
(bimb) and different metal ions for visible light-driven photocatalysis [47].
Degradation of the organic dye X3B was used to study the photocatalytic
performance of the MOFs, and three of them, [Co(btec)0.5(bimb)]n (6),
[Ni(btec)0.5(bimb)(H2O)]n (7), and [Cd(btec)0.5(bimb)0.5]n (8), exhibited better
performance than the commercial TiO2 (Degussa P-25) under visible light
(Fig. 7). The formation of hydroxyl radical OH∙ on the surface of the MOFs
under visible light was detected by capture with terephthalic acid to form
2-hydroxyterephthalic acid. By measuring the photoluminescence (PL) intensity
of the 2-hydroxyterephthalic acid, the authors were able to track the generation of
hydroxyl radical and a zero-order kinetic was observed (Fig. 8).
Metal-Organic Frameworks for Photocatalysis 97

Fig. 7 Degradation profiles of X3B under visible light irradiation in the presence of (a) without
catalyst, (b) Degussa P-25, (c) 6, (d ) 7, and (e) 8. Reprinted with permission from [47]. Copyright
2012 American Chemical Society

Fig. 8 (a) PL spectral changes observed during illumination of 8 in a 5  104 M basic solution
of terephthalic acid (excitation at 315 nm). (b) Comparison of the induced PL intensity at 425 nm
for MOFs 6 (red), 7 (blue), and 8 (black). Reprinted with permission from [47]. Copyright 2012
American Chemical Society

5 Photocatalytic Organic Transformations with MOFs

Compared to simple degradation of organic compounds, photocatalytic organic


transformations often require better control of the reaction and better selectivity
towards desired products, thus presenting a greater challenge for MOF
photocatalysis. In the last few years, small molecule dyes such as Ru(bpy)32+
and Ir(ppy)2(bpy)+ have been utilized as photoredox catalysts for various
organic reactions [20–24]. Given the high cost of these noble metal-based
98 T. Zhang and W. Lin

n+
CO2H CO2H

N
X N DMF, 100oC
ZrCl4 + + M [ Zr6(m3-O)4(m3-OH)4(bpdc)6-x(L)x ]
N N
X 9: L = L6
10: L = L7
CO2H CO2H
H2bpdc H2L6: M = Ir, X = C, n = 1
H2L7: M = Ru, X = N, n = 2

Scheme 2 Synthesis of phosphor-doped UiO-67 frameworks 9 and 10

Table 1 Photocatalytic aza-Henry reactions with 9 and 10

N 1 mol% catalyst, CH3NO2


N
air, fluorescent lamp (26 W), 12 h
H
X NO2 X
R1 X = H P1 X = H
R2 X = Br P2 X = Br
R3 X = CH3O P3 X = CH3O

Compounds Product 9 (%) 10 (%) Et2L6a (%) Et2L7a (%)


R1 P1 59 86 99 97
R2 P2 62 68 90 88
R3 P3 96 97 >99 >99
a
Diethyl esters of L6 and L7 (Et2L6 and Et2L7) were used as homogeneous controls

Table 2 Photocatalytic aerobic amine coupling reaction based on 10


NH2 N
1 mol% 10, CH3CN

X 300 W Xe lamp, 60oC, 1 h X X

R4 X = H P4 X = H

R5 X = CH3 P5 X = CH3

R6 X = OCH3 P6 X = OCH3

Compounds Product No light (%) No catalyst (%) Et2L7 (%) 10 (%) Reused 10 (%)
R4 P4 6 8 96 83 80
R5 P5 90
R6 P6 46

photocatalysts, it is desirable to develop heterogeneous version of these


photocatalysts to facilitate recycling and reuse. In 2011, Lin and coworkers
reported the preparation of highly stable photoactive UiO-67 MOFs 9 and 10
Metal-Organic Frameworks for Photocatalysis 99

Fig. 9 Schematic representation of mirror image structures of Zn-BCIP1 and Zn-BCIP2 and
their deprotected forms Zn-PYI1 or Zn-PYI2. Reprinted with permission from [56]. Copyright
2012 American Chemical Society

doped with Ru or Ir-based phosphors (Scheme 2) [35]. They investigated


aza-Henry reaction of tertiary amines, oxidative coupling of amines, and pho-
tooxidation of sulfides as the model reactions to illustrate the photocatalytic
activities of the MOFs 9 and 10. As shown in Table 1 and 2, 9 and 10
effectively catalyzed these reactions and gave slightly lower but comparable
yields of the desiring product as homogeneous controls. The MOFs exhibited
excellent reusability as no leaching of noble metals or decreasing of catalytic
100 T. Zhang and W. Lin

Table 3 Conversions and the ee values (in the parenthesis) of the photocatalytic α-alkylation of
aldehydesa
O Br
N O CO2Et
+ +
X EtO2C CO2Et fluorrescent lamp (26 W)
H * CO2Et
0.5 mol% catalyst, N2, 24 h
2 : 2 : 1 X

X= n-C6H13
, ,
Et
R7 R8 R9

Catalyst R7 R8 R9
Zn-PYI1 74 (+92) 65 (+86) 84 (+92)
Zn-PYI2 73 (81) 61 (78) 85 (89)
Ho-TCA/L-PYIa 86 (+20) 90 (+21) 93 (+20)
Ho-TCA/D-PYIa 85 (21) 90 (20) 95 (20)
MOF-150/L-PYIa 67 (+21) 78 (+24) 80 (+20)
MOF-150/D-PYIa 62 (22) 73 (22) 90 (21)
a
In the presence of additional chiral D/L-PYI (20 mol%)

activity was observed. Crystallinity of the MOFs was also retained after
photocatalysis. However, size selectivity was not observed in these reactions,
suggesting that it is possible that all of these reactions are mediated by
photochemically generated singlet oxygen.
Metalloporphyrin-based MOFs have also been investigated for light-harvesting
and photochemical properties [53, 54]. In 2011, Wu et al. reported the synthesis of a
tin-porphyrin-derived MOF [Zn2(H2O)4SnIV(TPyP)(HCOO)2]·4NO3·DMF·4H2O
(11, SnIVTPyP ¼ 5,10,15,20-tetra(4-pyridyl)-tin(IV)-porphyrin) [55]. 11
showed high photocatalytic performance for the selective oxidation of 1,5-
dihydroxynaphehalene to 5-hydroxynaphthalene-1,4-dione. The authors also tested
the photocatalytic activity of 11 with the oxidation of sulfides into sulfoxides and
found that 11 exhibited better selectivity towards sulfoxides versus sulfones than
SnIV(OH)2TPyP.
Catalytically active chiral MOFs are of much interest since they provide an
effective way to immobilize chiral active sites into a solid, making it possible to
prepare heterogeneous asymmetric catalysts. Combination of an asymmetric
catalytic site and a photoactive functional component in the same MOF can
potentially lead to highly efficient and stereoselective photocatalysis. Applying
this strategy, Duan et al. reported in 2012 the incorporation of an asymmetric
organocatalyst L or D-pyrrolidine-2-yl-imidazole (PY1) into a photoactive MOF
with a triphenylamine moiety (Fig. 9) [56]. The resulting MOFs, Zn-PYI1 or
Zn-PYI2, were then used as heterogeneous catalysts for light-driven asymmetric
α-alkylation of aliphatic aldehydes (Table 3). The photocatalytic reactions gave
high yields and remarkable enantioselectivities. MOFs with different chiralities
gave products with opposite selectivities. Control experiments using the Ho-TCA
Metal-Organic Frameworks for Photocatalysis 101

(TCA ¼ tricarboxytriphenylamine) MOF and MOF-150, which bear the same


photoactive ligand, as photocatalyst with the same chiral adduct gave much
lower ee value though comparable yields were achieved. The results indicated
that the combination of the photosensitizer and the asymmetric organocatalyst in
the same framework is the key to an enantioselective and photoactive MOF
catalyst.

6 Conclusions and Outlook

Because photocatalytic systems typically require multiple components, MOFs are


potentially advantageous in photocatalysis due to the ability to incorporate multiple
functional groups in MOFs with well-defined spatial arrangements. In the past
several years, MOFs have shown promising applications on photocatalysis with
impressive results on photochemical hydrogen evolution, CO2 reduction, and
organic reactions. However, much more work is needed to prepare practically
useful photocatalytic MOF or MOF-based assemblies.
Many MOFs tend to have modest hydrolytic stability which greatly limits their
applications in solar energy utilization. Photoactive cross-linked polymers [57–59]
could be an alternative choice for designing new photocatalytic systems. As MOFs
and cross-linked polymers have shown great possibilities for multifunctiona-
lization, further development in these areas can lead to novel materials with
multicomponents and hierarchical structures for photocatalysis.

Acknowledgment We thank NSF (DMR-1308229) for funding support.

References

1. Fujishima A, Honda K (1972) Electrochemical photolysis of water at a semiconductor


electrode. Nature 238:37–38
2. Youngblood WJ, Lee SA, Kobayashi Y, Hernandez-Pagan E, Hoertz PG, Moore TA, Moore AL,
Gust D, Mallouk TE (2009) Photoassisted overall water splitting in a visible light-absorbing
dye-sensitized photoelectrochemical cell. J Am Chem Soc 131:926–927
3. Reece SY, Hamel JA, Sung K, Jarvi TD, Esswein AJ, Pijpers JJH, Nocera DG (2011) Wireless
solar water splitting using silicon-based semiconductors and earth-abundant catalysts. Science
334:645–648
4. Ohno T, Bai L, Hisatomi T, Maeda K, Domen K (2012) Photocatalytic water splitting using
modified GaN:ZnO Solid solution under visible light: long-time operation and regeneration of
activity. J Am Chem Soc 134:8254–8259
5. Wang X, Xu Q, Li M, Shen S, Wang X, Wang Y, Feng Z, Shi J, Han H, Li C (2012)
Photocatalytic overall water splitting promoted by an phase junction on Ga2O3. Angew
Chem Int Ed 51:13089–13092
6. Teets TS, Nocera DG (2011) Photocatalytic hydrogen production. Chem Commun
47:9268–9274
102 T. Zhang and W. Lin

7. Du P, Eisenberg R (2012) Catalysts made of earth-abundant elements (Co, Ni, Fe) for water
splitting: recent progress and future challenges. Energy Environ Sci 5:6012–6021
8. Lakadamyali F, Kato M, Muresan NM, Reisner E (2012) Selective reduction of aqueous
protons to hydrogen with a synthetic cobaloxime catalyst in the presence of atmospheric
oxygen. Angew Chem Int Ed 51:9381–9384
9. Luo S, Mejı́a E, Friedrich A, Pazidis A, Junge H, Surkus A, Jackstell R, Denurra S, Gladiali S,
Lochbrunner S, Beller M (2013) Photocatalytic water reduction with copper-based photosen-
sitizers: a noble-metal-free system. Angew Chem Int Ed 52:419–423
10. Hetterscheid DGH, Reek JNH (2012) Mononuclear water oxidation catalysts. Angew Chem
Int Ed 51:9740–9747
11. Duan L, Bozoglian F, Mandal S, Stewart B, Privalov T, Llobet A, Sun L (2012) A molecular
ruthenium catalyst with water-oxidation activity comparable to that of photosystem II. Nat
Chem 4:418–423
12. Rosi NL, Eckert J, Eddaoudi M, Vodak DT, Kim J, O’Keeffe M, Yaghi OM (2003) Hydrogen
storage in microporous metal-organic frameworks. Science 300:1127–1129
13. Murray LJ, Dinca M, Long JR (2009) Hydrogen storage in metal-organic frameworks. Chem
Soc Rev 38:1294–1314
14. Lee J, Farha OK, Roberts J, Scheidt KA, Nguyen ST, Hupp JT (2009) Metal-organic frame-
work materials as catalysts. Chem Soc Rev 38:1450–1459
15. Ma L, Abney C, Lin W (2009) Enantioselective catalysis with homochiral metal-organic
frameworks. Chem Soc Rev 38:1248–1256
16. Della Rocca J, Liu D, Lin W (2011) Nanoscale metal-organic frameworks for biomedical
imaging and drug delivery. Acc Chem Res 44:957–968
17. Horcajada P, Gref R, Baati T, Allan PK, Maurin G, Couvreur P, Férey G, Morris RE, Serre C
(2012) Metal-organic frameworks in biomedicine. Chem Rev 112:1232–1268
18. Li J, Sculley J, Zhou H (2012) Metal-organic frameworks for separations. Chem Rev
112:869–932
19. Kreno LE, Leong K, Farha OK, Allendorf M, Van Duyne RP, Hupp JT (2012) Metal-organic
framework materials as chemical sensors. Chem Rev 112:1105–1125
20. Nicewicz DA, MacMillan DWC (2008) Merging photoredox catalysis with organocatalysis:
the direct asymmetric alkylation of aldehydes. Science 322:77–80
21. Nagib DA, Scott ME, MacMillan DWC (2009) Enantioselective α-trifluoromethylation of
aldehydes via photoredox organocatalysis. J Am Chem Soc 131:10875–10877
22. Du J, Yoon TP (2009) Crossed intermolecular [2+2] cycloadditions of acyclic enones via
visible light photocatalysis. J Am Chem Soc 131:14604–14605
23. Condie AG, González-Gómez JC, Stephenson CRJ (2010) Visible-light photoredox catalysis:
Aza-Henry reactions via CH functionalization. J Am Chem Soc 132:1464–1465
24. Lang X, Ji H, Chen C, Ma W, Zhao J (2011) Selective formation of imines by aerobic
photocatalytic oxidation of amines on TiO2. Angew Chem Int Ed 50:3934–3937
25. Kataoka Y, Sato K, Miyazaki Y, Masuda K, Tanaka H, Naito S, Mori W (2009) Photocatalytic
hydrogen production from water using porous material [Ru2(p-BDC)2]n. Energy Environ Sci
2:397–400
26. Cavka JH, Jakobsen S, Olsbye U, Guillou N, Lamberti C, Bordiga S, Lillerud KP (2008) A new
zirconium inorganic building brick forming metal organic frameworks with exceptional
stability. J Am Chem Soc 130:13850–13851
27. Gomes Silva C, Luz I, Llabrés i Xamena FX, Corma A, Garcı́a H (2010) Water stable
Zr–benzenedicarboxylate metal-organic frameworks as photocatalysts for hydrogen genera-
tion. Chem Eur J 16:11133–11138
28. Wang C, deKrafft KE, Lin W (2012) Pt nanoparticles@photoactive metal-organic
frameworks: efficient hydrogen evolution via synergistic photoexcitation and electron injec-
tion. J Am Chem Soc 134:7211–7214
Metal-Organic Frameworks for Photocatalysis 103

29. Fateeva A, Chater PA, Ireland CP, Tahir AA, Khimyak YZ, Wiper PV, Darwent JR,
Rosseinsky MJ (2012) A water-stable porphyrin-based metal-organic framework active for
visible-light photocatalysis. Angew Chem Int Ed 51:7440–7444
30. Jordan P, Fromme P, Witt HT, Klukas O, Saenger W, Krausz N (2001) Three-dimensional
structure of cyanobacterial photosystem I at 2.5 Å resolution. Nature 411:909–917
31. Hawecker J, Lehn J, Ziessel R (1983)Efficient photochemical reduction of CO2 to CO by
visible light irradiation of systems containing Re(bipy)(CO)3X or Ru(bipy)32+-Co2+
combinations as homogeneous catalysts. J Chem Soc Chem Commun 0:536–538
32. Lin W, Frei H (2005) Photochemical CO2 splitting by metal-to-metal charge-transfer excita-
tion in mesoporous ZrCu(I)-MCM-41 silicate sieve. J Am Chem Soc 127:1610–1611
33. Liu Q, Zhou Y, Kou J, Chen X, Tian Z, Gao J, Yan S, Zou Z (2010) High-yield synthesis of
ultralong and ultrathin Zn2GeO4 nanoribbons toward improved photocatalytic reduction of
CO2 into renewable hydrocarbon fuel. J Am Chem Soc 132:14385–14387
34. Yan S, Ouyang S, Gao J, Yang M, Feng J, Fan X, Wan L, Li Z, Ye J, Zhou Y, Zou Z (2010) A
room-temperature reactive-template route to mesoporous ZnGa2O4 with improved
photocatalytic activity in reduction of CO2. Angew Chem Int Ed 49:6400–6404
35. Wang C, Xie Z, deKrafft KE, Lin W (2011) Doping metal-organic frameworks for water
oxidation, carbon dioxide reduction, and organic photocatalysis. J Am Chem Soc
133:13445–13454
36. Koike K, Hori H, Ishizuka M, Westwell JR, Takeuchi K, Ibusuki T, Enjouji K, Konno H,
Sakamoto K, Ishitani O (1997) Key process of the photocatalytic reduction of CO2 using
[Re(4,40 -X2-bipyridine)(CO)3PR3]+ (X ¼ CH3, H, CF3; PR3 ¼ Phosphorus Ligands): dark
reaction of the one-electron-reduced complexes with CO2. Organometallics 16:5724–5729
37. Takeda H, Koike K, Inoue H, Ishitani O (2008) Development of an efficient photocatalytic
system for CO2 reduction using rhenium(I) complexes based on mechanistic studies. J Am
Chem Soc 130:2023–2031
38. Fu Y, Sun D, Chen Y, Huang R, Ding Z, Fu X, Li Z (2012) An amine-functionalized titanium
metal-organic framework photocatalyst with visible-light-induced activity for CO2 reduction.
Angew Chem Int Ed 51:3364–3367
39. Llabrés IX, Corma A, Garcia H (2007) Applications for metalorganic frameworks (MOFs) as
quantum dot semiconductors. J Phys Chem C 111:80–85
40. Alvaro M, Carbonell E, Ferrer B, Llabrési Xamena F, Garcia H (2007) Semiconductor
behavior of a metal-organic framework (MOF) Chem Eur J 13:5106–5112
41. Wen L, Wang F, Feng J, Lv K, Wang C, Li D (2009) Structures, photoluminescence, and
photocatalytic properties of six new metalorganic frameworks based on aromatic
polycarboxylate acids and rigid imidazole-based synthons. Cryst Growth Des 9:3581–3589
42. Wang X, Wang Y, Liu G, Tian A, Zhang J, Lin H (2011) Novel inorganic-organic hybrids
constructed from multinuclear copper cluster and Keggin polyanions: from 1D wave-like
chain to 2D network. Dalton Trans 40:9299–9305
43. Wang F, Ke X, Zhao J, Deng K, Leng X, Tian Z, Wen L, Li D (2011) Six new metal-organic
frameworks with multi-carboxylic acids and imidazole-based spacers: syntheses, structures
and properties. Dalton Trans 40:11856–11865
44. Du J, Yuan Y, Sun J, Peng F, Jiang X, Qiu L, Xie A, Shen Y, Zhu J (2011) New photocatalysts
based on MIL-53 metal-organic frameworks for the decolorization of methylene blue dye. J
Hazard Mater 190:945–951
45. Das MC, Xu H, Wang Z, Srinivas G, Zhou W, Yue Y, Nesterov VN, Qian G, Chen B (2011) A
Zn4O-containing doubly interpenetrated porous metal-organic framework for photocatalytic
decomposition of methyl orange. Chem Commun 47:11715–11717
46. Zhou S, Kong Z, Wang Q, Li C (2012) Synthesis, structure and photocatalytic property of a
novel 3D (3,8)-connected metal-organic framework based on a flexible triphosphonate and a
pentanuclear Cu(II) unit. Inorg Chem Commun 25:1–4
104 T. Zhang and W. Lin

47. Wen L, Zhao J, Lv K, Wu Y, Deng K, Leng X, Li D (2012) Visible-light-driven photocatalysts


of metal-organic frameworks derived from multi-carboxylic acid and imidazole-based spacer.
Cryst Growth Des 12:1603–1612
48. Kan W, Liu B, Yang J, Liu Y, Ma J (2012) A series of highly connected metal-organic
frameworks based on triangular ligands and d10 metals: syntheses, structures, photolumi-
nescence, and photocatalysis. Cryst Growth Des 12:2288–2298
49. Guo J, Yang J, Liu Y, Ma J (2012) Two novel 3D metal-organic frameworks based on two
tetrahedral ligands: syntheses, structures, photoluminescence and photocatalytic properties.
CrystEngComm 14:6609–6617
50. Zhang Z, Yang J, Liu Y, Ma J (2013) Five polyoxometalate-based inorganic–organic hybrid
compounds constructed by a multidentate N-donor ligand: syntheses, structures, electrochem-
istry, and photocatalysis properties. CrystEngComm 15:3843–3853
51. Du P, Yang Y, Yang J, Liu B, Ma J (2013) Syntheses, structures, photoluminescence,
photocatalysis, and photoelectronic effects of 3D mixed high-connected metal-organic
frameworks based on octanuclear and dodecanuclear secondary building units. Dalton Trans
42:1567–1580
52. Li H, Eddaoudi M, O’Keeffe M, Yaghi OM (1999) Design and synthesis of an exceptionally
stable and highly porous metal-organic framework. Nature 402:276–279
53. Son H, Jin S, Patwardhan S, Wezenberg SJ, Jeong NC, So M, Wilmer CE, Sarjeant AA,
Schatz GC, Snurr RQ, Farha OK, Wiederrecht GP, Hupp JT (2013) Light-harvesting and
ultrafast energy migration in porphyrin-based metal-organic frameworks. J Am Chem Soc
135:862–869
54. Lee CY, Farha OK, Hong BJ, Sarjeant AA, Nguyen ST, Hupp JT (2011) Light-harvesting
metal-organic frameworks (MOFs): efficient strut-to-strut energy transfer in bodipy and
porphyrin-based MOFs. J Am Chem Soc 133:15858–15861
55. Xie M, Yang X, Zou C, Wu C (2011) A SnIV–porphyrin-based metal-organic framework for
the selective photo-oxygenation of phenol and sulfides. Inorg Chem 50:5318–5320
56. Wu P, He C, Wang J, Peng X, Li X, An Y, Duan C (2012) Photoactive chiral metal-organic
frameworks for light-driven asymmetric α-alkylation of aldehydes. J Am Chem Soc
134:14991–14999
57. Xie Z, Wang C, deKrafft KE, Lin W (2011) Highly stable and porous cross-linked polymers
for efficient photocatalysis. J Am Chem Soc 133:2056–2059
58. Wang C, Xie Z, deKrafft KE, Lin W (2012) Light-harvesting cross-linked polymers for
efficient heterogeneous photpcatalysis. ACS Appl Mater Interfaces 4:2288–2294
59. Wang J, Wang C, deKrafft KE, Lin W (2012) Cross-linked polymers with exceptionally high
Ru(bipy)32+ loadings for efficient heterogeneous photocatalysis. ACS Catal 2:417–424
Struct Bond (2014) 157: 105–144
DOI: 10.1007/430_2013_121
# Springer-Verlag Berlin Heidelberg 2013
Published online: 9 August 2013

Metal-Organic Frameworks for


Photochemical Reactions

Raghavender Medishetty and Jagadese J. Vittal

Abstract In this chapter, various photochemical reactions involving coordina-


tion polymers and metal-organic frameworks have been described. These include
structural transformations, post-synthetic modification of surfaces to create reactive
sites, radicals which are not possible by convention synthetic routes, polymeriza-
tion on the surfaces, and cis–trans isomerization of the guest molecules in the
cavities using the dynamic behavior of the structures. Further, a few examples
have been provided where [2þ2] cycloaddition reaction has been used as a
tool to monitor the changes in the solid-state structures in the absence of crystal
structures. The structures of the resultant products of a number of polymerization
reactions such as alkenoates were hampered due to the lack of suitable structural
tool other than single crystal X-ray crystallography. On the other hand, most of
the examples discussed in this chapter do maintain the single crystals at the end
of the reactions making them amenable for structural elucidation to understand
the reactivity completely.

Keywords 2þ2 cycloaddition  Coordination polymers (CPs)  Crystal engineering 


Cyclobutane derivatives  Ladder structures  MOFs  Photo-reactivity  Solid-state
reactions  Structural transformation

Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
2 Polymerization of Metal Complexes by [2þ2] Cycloaddition Reactions . . . . . . . . . . . . . . . 107
2.1 Role of Ag  C Interactions in Transferring a Dimer to a 1D CP . . . . . . . . . . . . . . . . . 108
2.2 Polymerization of a Dimer Through [2þ2] Cycloaddition Reaction . . . . . . . . . . . . . . 109
2.3 Photo-Reactivity of 1D H-Bonded Zwitter-Ionic Metal Complexes . . . . . . . . . . . . . . . 110
2.4 Polymerization of a Hydrogen-Bonded Metal Complex Containing Several
Parallel and Crisscross C═C Bonds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

R. Medishetty and J.J. Vittal (*)


Department of Chemistry, National University of Singapore, Singapore 117543, Singapore
e-mail: chmjjv@nus.edu.sg
106 R. Medishetty and J.J. Vittal

3 Photo-Reactivity in 1D Polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113


3.1 1D to 1D0 Transformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
3.2 Metal Template to Align the C═C Bonds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
3.3 Alignment of the Spacer Ligand bpe in an Infinitely Parallel Fashion . . . . . . . . . . . . 116
3.4 Stepwise Double [2þ2] Cycloaddition Reaction Assisted by Ag  Ag Interactions 117
3.5 The Use of Conjugated Dienes for Double [2þ2] Cycloaddition Reactions in 1D
Polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
3.6 Construction of Photo-Reactive 1D CP Using a Bioactive Dicarboxylates . . . . . . . 118
4 1D to 2D Transformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
4.1 Synthesis of Unique Furan-Based Ligand . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
4.2 Photo-Reactivity of a Triple-Stranded CP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
5 1D to 3D Transformation in CPs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
5.1 An Unusual Orientation of bpe Pairs for Photo-Reactivity . . . . . . . . . . . . . . . . . . . . . . . . 122
6 Photo-Reactivity in 2D CPs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
6.1 Photo-Reactivity of a Layer CP yielding an unusual product . . . . . . . . . . . . . . . . . . . . . . 122
6.2 2D to 2D0 Transformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
6.3 2D to 2D0 Transformation and Its Thermal Cyclo-Reversion . . . . . . . . . . . . . . . . . . . . . . 124
6.4 Influence of 2D to 2D0 Structural Conversion on the Gas Sorption Properties . . . 125
6.5 2D to 2D0 and 3D to 3D0 Structural Transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
6.6 2D to 3D Transformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
6.7 Transformation of 2D Interdigitated CP to 3D MOF . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
6.8 Post-Synthetic Modification to Fine-Tune the Sorption Behavior of 2D Sheets . . 129
7 Photo-Reactive 3D CPs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
7.1 Post-Synthetic Modification of Double-Pillared MOFs by [2þ2] Cycloaddition
Reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
7.2 Photo-Reactive Double-Pillared MOFs Containing Conjugated Olefin Bonds . . . . 131
7.3 Host–Guest Reactions in Pillared-Layer MOFs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
7.4 Cis–Trans Isomerization on the Gas Adsorption Property of MOFs . . . . . . . . . . . . . . 134
7.5 Photo-Reactive Lanthanide-MOFs (Ln-MOF) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
8 [2þ2] Cycloaddition Reaction to Monitor the Structural Transformations in CPs . . . . . 137
8.1 Anisotropic Movements of 1D CPs by Desolvation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
8.2 Transformation of a Linear CP to a Ladder Structure by Thermal Dehydration . . 137
8.3 Loss of Water Chains Transforms a Linear CP to a Ladder Structure . . . . . . . . . . . . 139
9 Photopolymerization of Alkenoate Systems in CPs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
10 Photo-Reactivity in Solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
11 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

1 Introduction

Among the solid-state photochemical reactions, [2þ2] cycloaddition reaction


has been studied extensively in the past several decades [1–11]. The classical
work of G. Schmidt and coworkers in 1960s paved away the field of solid-
state organic photochemistry as well as crystal engineering [12–14]. Due to this
historical importance, this double bonds containing spacer and terminal ligands
have been incorporated into coordination polymers (CPs) and porous CPs (PCPs)
which is also popularly called as metal-organic frameworks (MOFs) to design
Metal-Organic Frameworks for Photochemical Reactions 107

photo-reactive solids [15–17]. The metal-coordination bond offers very important


role in directing the organic linkers and/or guest molecules to satisfy the geometry
criteria required for the photoreactions. At the initial stage, enormous interests
were shown to utilize various crystal engineering tools to design photo-reactive
CPs and MOFs. Since the crystallization process is controlled by the kinetic factors,
the challenge to design photo-reactive CPs still persists [18–20]. However, the
interest is slowly shifting from design strategies to exploration and understanding
the structure–function–property relationships. This includes molecular sensing,
controlled guest release, pore size, and surface functionalization.
In this chapter, the following will be addressed: (1) How the olefin pairs in
the spacer ligands as well as guest molecules can be aligned to undergo [2þ2]
cycloaddition reactions. (2) In the photochemical reactions, new σ bonds are
formed between two carbon atoms to yield cyclobutane rings. This process
brings structural transformations due to change in the structures, connectivity,
dimensionality, or interpenetration. Hence, sections are mainly divided based on
the changes in the dimensionality by the formation of cyclobutane rings from the
olefin pairs. (3) The changes in the physical and chemical properties with the
structural transformations will also be discussed. (4) Apart from solid-state [2þ2]
cycloaddition reactions, other photochemical reactions such as polymerizations
and photodissociations have also been discussed. (5) Photochemical reactions in
solution have also been briefly mentioned for the sake of completion.

2 Polymerization of Metal Complexes by [2þ2]


Cycloaddition Reactions

The first [2þ2] photopolymerization reaction of a metal complex was reported for
the mononuclear complex of uranyl chloride with trans,trans-dibenzylideneacetone
(dba) shown in Fig. 1 resulted in a stereospecific rctt isomer of truxillic type 1,3-bis
(3-phrnylprop-2-enol-oxo-2,4-bis(phenyl)cyclobutane) by Praetorius and Korn in
1910 [21, 22]. Later, Stobbe and Farber reported the formation of the similar
dimer using SnCl4 in 1925 [23] with a closely related ligand. To understand the
reactivity, Alcock determined the crystal structure of [(UO2)Cl2(dba)2] (1) and
[(UO2)Cl2(dba)2]·2CH3COOH (2), in which the uranium is present in octahedral
coordination geometry surrounded by two Cl atoms, two O atoms of uranyl group,
and two keto oxygen atoms from dba [24]. The ligand dba has trans,trans confor-
mation, and the olefinic groups are separated by 4.09 and 3.51 Å for solvated
and desolvated forms, respectively. Later, the complexes of Ni(II), Cu(II), and
Zn(II) with a series of substituted cyclopentanones were found to undergo [2þ2]
photo-cycloaddition reactions [25] (Fig. 2).
108 R. Medishetty and J.J. Vittal

Fig. 1 The alignment of olefin groups of dba ligand in a uranyl chloride complex

Fig. 2 The alignment of olefin groups of dba ligand in the SnCl4 complex

2.1 Role of Ag  C Interactions in Transferring


a Dimer to a 1D CP

MacGillivray and coworkers showed that the intermolecular Ag  C bond formed
during [2þ2] cycloaddition reaction in a discrete dinuclear Ag(I) complex can be
used in the formation of an interesting 1D CP [26]. The 4spy ligands in the metal
complex [Ag2(4spy)4][TFA]2 (3) (4spy ¼ 4-stryrylpyridine) have been paired up
with the help of argentophilic and π–π interactions [26]. Though the C═C bonds in
4spy ligand pairs are aligned in a crisscross manner, the pedal motion of C═C bonds
assisted the quantitative photo-reactivity of this compound and resulted in the
formation of [Ag2(rctt-4-ppcb)2][(TFA)2] (4) (rcct-4-ppcb ¼ 1,3-bis(40 -pyridyl)-
2,4-bis(phenyl)cyclobutane). This pedal motion was well studied by Ogawa and
coworkers for stilbenes and azo-stilbenes [27]. Interestingly, this transformation
Metal-Organic Frameworks for Photochemical Reactions 109

Fig. 3 Transformation of a metal complex into a 1D CP due to the formation of Ag  C interactions

took place in an SCSC manner without loss of single crystallinity that enabled
the structural elucidation of final dimerized compound by X-ray crystallography.
The solid-state structure of the photoproduct showed that significant movements
of the Ag(I) atoms and rctt-4-ppcb ligands took place during photoreaction
which resulted in the formation of Ag  C interaction (distance 2.64 Å) between
the neighboring molecules giving raise to the formation of a 1D CP as shown
in Fig. 3.

2.2 Polymerization of a Dimer Through [2þ2] Cycloaddition


Reaction

Recently another photoactive metal complex, [Ag2(Cl-spy)4(ClO3)2] (5) (where


Cl-spy ¼ trans-1-(4-Cl-3-pyridyl)-2-(phenyl)ethylene), was synthesized with Ag(I)
in tetrahedral coordination geometry and coordinated by N of two Cl-spy, O atoms
of two ClO3 anions [28]. The solid-state structure analysis showed that only 50%
of Cl-spy are aligned parallel between the dinuclear complexes and are separated by
3.73 Å as shown in Fig. 4. The other 50% are arranged beyond Schmidt’s criteria
(5.93 and 6.29 Å), and the two rings in the Cl-spy are twisted from coplanarity
by 20.6 . As expected, on UV irradiation of this compound showed photo-reactivity
of this compound in an SCSC manner and resulted in the formation of 1D CP.
110 R. Medishetty and J.J. Vittal

Fig. 4 Polymerization of a metal complex into a 1D CP through [2+2] cycloaddition reaction

2.3 Photo-Reactivity of 1D H-Bonded Zwitter-Ionic


Metal Complexes

H-bonds are one of the strongest directional forces among all the supramolecular
interactions. In a zwitter-ionic metal complex [Pb(bpe-H)2(TFA)4] (6) (where
TFA ¼ trifluoroacetate anion), complementary hydrogen-bonds between a coordi-
nated monoprotonated bpe-Hþ cation and the neighboring oxygen atom of the TFA
ligand assists the alignment of the bpe-Hþ pairs as shown in Fig. 5. The Pb(II)
center is coordinated by four trifluoroacetates and two bpe-Hþ cations. The infi-
nitely organized coordinated bpe-Hþ pairs easily undergo photocyclization under
UV irradiation quantitatively to give the anticipated rctt-tpcb isomer [29]. The
colorless block-like crystals turned into gel in 1 h and became powder after 25 h
of UV irradiation. During photodimerization, two CF3CO2H were eliminated due
to the proton transfer from bpe-Hþ to TFA ligand bonded to the Pb(II), resulted
in [Pb(rctt-tpcb)(TFA)2] (7). Since this is not an SCSC reaction, the structure
of the photoproduct was determined after recrystallizing from methanol. Due to
the reorganization of structures during crystallization, the connectivity of the
crystallized product of [Pb(rctt-tpcb)(TFA)2] is slightly different from predicted
structure (shown in Fig. 5) based on the connectivity of the precursor.
A very similar zwitter-ionic complex, [Cd(bpe-H)2(H2O)(S2O3)2] (8), was
reported by Natarajan et al., in which two protonated [bpe-H]þ ligands are coordi-
nated to Cd(II) in trans fashion [30]. The pyridinium N–H proton at the other
Metal-Organic Frameworks for Photochemical Reactions 111

Fig. 5 A zigzag hydrogen-bonded zwitter-ionic Pb(II) complex undergoes polymerization reaction


under UV light

end of [bpe-H]þ is hydrogen bonded to the neighboring thiosulfate O atom to


form a hydrogen-bonded 1D polymer. The olefinic groups are separated by 3.7 Å
and upon UV irradiation of this compound showed quantitative [2þ2] photo-
cycloaddition as confirmed by 1H-NMR and resulted in the formation of 1D chains
as shown in Fig. 6. However, the solid-state structure of final compound was not
characterized.

2.4 Polymerization of a Hydrogen-Bonded Metal


Complex Containing Several Parallel and
Crisscross C═C Bonds

An interesting H-bonded 1D metal complex, [Zn(bpe)2(H2O)4](NO3)2·(8/3)


H2O·(2/3)bpe (9), was reported, where the Zn(II) is present in octahedral geometry
arising from the coordination of two bpe ligands in trans fashion and four water
112 R. Medishetty and J.J. Vittal

Fig. 6 Polymerization of a metal complex to a 1D CP by [2+2] cycloaddition reaction

Fig. 7 A diagram showing the alignment of bpe molecules in the hydrogen-bonded metal
complex, 9

molecules in the equatorial sites [31]. A few guest bpe molecules were also found in
the lattice along with lattice water. Only one side of the bpe is coordinated to
Zn(II), and the other side of bpe is H-bonded to coordinated water and forming 1D
H-bonded metal complex. These complementary H-bonds drive the bpe molecules
to align parallel with each other. Interestingly, this compound consists of totally six
different types of bpe molecules that are aligned in parallel as shown in Fig. 7.
However, the C═C bonds are aligned both in parallel and crisscross manner and
are separated 3.55–3.88 Å from each other. On UV irradiation of the single crystals
for 25 h showed 46% yield tpcb comprising 39% rctt and 7% rtct isomers.
But this percentage of rtct isomer formed is lower than expected percentage
based on the number of antiparallel orientation of C═C bonds. This has been
attributed to the partial pedal motion of the C═C bonds. This pedal motion in
this compound has been found to be accelerated by the mechanical grinding by
pestle and mortar for 5 min along with increased the photo-reactivity to 100%
tpcb when exposed under UV light for 25 h. In the absence of the crystal structure,
the resultant product is expected to be a 1D CP similar to those described in
Sect. 2.3. Hence, this reaction can be considered as metal complex polymerization.
Metal-Organic Frameworks for Photochemical Reactions 113

3 Photo-Reactivity in 1D Polymers

1D chains are the simplest of all the CPs. Of these, molecular ladder polymers
dictated by the proximity of two or more metal ions, spacer ligands, and anions
have attracted attention due to their interesting magnetic electronic and optical
properties. Furthermore, among the 1D CPs only the ladder structure is useful to
carry out [2þ2] cycloaddition reaction. However, the synthesis of 1D photo-
reactive CPs is often challenging, due to the formation of other kinetic products
during the crystallization process.

3.1 1D to 1D0 Transformation

In 2005, Vittal’s laboratory reported the synthesis of a ladder-like photo-reactive


CP, [{(H3CCO2)(μ-O2CCH3)Zn}2(μ-bpe)2] (10), in which the Zn(II) atoms of
the two linear Zn(bpe) strands are bridged by two acetate ligands and chelated
by another acetate ligand. Here, the bridging acetate ligands act as “rungs” in the
resultant ladder structure and control the distance between the two linear Zn(bpe)
strands. As a result, the bpe ligands are aligned in parallel, and the olefinic groups
are separated by 3.66 Å. UV irradiation of this compound showed quantitative
photoreaction. However, the single crystals were not preserved in the photoproduct.
In contrast, in another similar ladder compound [{(TFA)(μ-O2CCH3)Zn}2(μ-bpe)2]
(11) in which half of the acetate groups were replaced by TFA, the single crystals
were intact at the end of the UV experiment. Therefore, the structure of the
photoproduct, [{(TFA)(μ-O2CCH3)Zn}2(rctt-tpcb)2] (12) (rctt-tpcb ¼ regio cis,
trans, and trans tetrakis(4-pyridyl)cyclobutane), was able to be determined by
X-ray crystallography (Fig. 8). During the photoreaction, the fluorine atoms
might have assisted to withstand the strain in the single crystals during the structural
transformation [32]. It is also worth noted that this is the first report on the [2þ2]
cycloaddition reaction observed in CPs in SCSC fashion.
These ladder polymers have also been obtained from mechanochemical synthesis
by grinding the stoichiometric ratios of the precursors as shown in Fig. 9 [33].
Recently, it was reported that benzoate ligands can also be used as bridging
ligands to bring a pair of linear [Pb(bpe)] linear strands to form photo-reactive ladder
polymer [34]. In the compounds [Pb2(μ-bpe)2(μ-O2C–C6H5)2(O2C–C6H5)2] (13),
[Pb2(μ-bpe)2(μ-O2C-p-Tol)2(O2C-p-Tol)2]·1.5H2O (14), and [Pb2(μ-bpe)2(μ-O2C-m-Tol)2
(O2C-m-Tol)2] (15), the bpe pairs have parallel orientations and found to undergo
quantitative [2þ2] cycloaddition reaction. Of these, the later compound with
m-Tol ligand has been found to undergo SCSC transformation to [Pb2(rctt-tpcb)
(μ-O2C-m-Tol)2(O2C-m-Tol)2] (16).
Lang and coworkers also succeeded in extending this design principle to
synthesize a photo-reactive Cd(II) ID CP with ladder structure [Cd2(μ-bpe)2
(μ-O2C–C6H4-p-Cl)2(O2C–C6H4-p-Cl)2] (17) using p-chlorobenzoate as the clipping
114 R. Medishetty and J.J. Vittal

Fig. 8 1D to 1D0 structural transformation using an infinitely parallel bpe pairs

Fig. 9 Schematic representation of mechanochemical synthesis of photo-reactive 1D CPs

agent for Cd(bpe) linear polymers. The quantitative photodimerization reaction was
also accompanied by SCSC structural transformation. The Cd  Cd distance
increased from 3.87 to 4.09 Å during the transformation as expected [35].
Metal-Organic Frameworks for Photochemical Reactions 115

Fig. 10 Schematic representation of photoreaction of an infinitely aligned bpe pairs in a 1D Schiff


base CP

3.2 Metal Template to Align the C═C Bonds

MacGillivray and coworkers have employed a template method to align a pair of


bpe molecules using a dinuclear Zn(II) Schiff base complex in which the distance
between the Zn(II) atoms is fixed at 3.1 Å. This Schiff base is able to coordinate
and occupy all the equatorial positions of the two Zn(II) atoms in octahedral
coordination geometry. Hence, the axial positions are freely available for the linear
spacer ligands. First, they have shown that a pair of bpe can be aligned using this
methodology. One of the two axial positions at the Zn(II) atoms was employed to
synthesize a discrete tetranuclear compound which undergoes [2þ2] cycloaddition
reaction in an SCSC manner [36]. Later, they have extended the same template
method successfully to construct a photo-reactive ladder structure. In this ladder
structure, both the axial positions of the Zn(II) atoms were utilized, and as a result,
the bpe pairs were aligned on both sides of the template as shown in Fig. 10.
UV irradiation of this compound showed quantitative cycloaddition reaction
resulting in another 1D ladder structure. However, single crystals could not survive
after the photoreaction, probably attributed to the loss of water molecules during
this photoreaction [37].
116 R. Medishetty and J.J. Vittal

Fig. 11 Schematic representation of the structural transformation of 18–19 by photodimerization


reaction

3.3 Alignment of the Spacer Ligand bpe in an Infinitely


Parallel Fashion

Tong et al. in 2011 have shown that it is possible to align the bpe pairs one below
the other infinitely in a 1D CP where the bpe ligands act as rungs [38]. Such rare
ladder structure has been observed in [Zn2(1,2-bdc)2(bpe)]·MeOH (18) (where
1,2-bdc is benzene-1,2-dicarboxylate). The nature of bonding of 1,2-bdc ligands
dictates the intrachain Zn  Zn separations of 3.48 and 4.45 Å. One of the carbox-
ylate group in one 1,2-bdc ligand chelates a Zn(II) in μ2:η1:η1 fashion, and the other
carboxylate group bridges two more Zn(II) atoms in μ2:η0:η2 fashion in the 1D
chain. On the other hand, each carboxylate group in the second 1,2-bdc bridges to
two different Zn(II) atoms but overall bonded to three different Zn(II) atoms. Two
adjacent 1D strands are bridged by bpe rungs to form this unique ladder structure as
shown in Fig. 11. The nonbonding distances between the double bonds of the
adjacent bpe ligands 3.84 and 3.96 Å satisfy Schmidt’s topochemical criteria.
Single crystal of this compound upon UV irradiation under a Hg lamp (400 W)
for 5 h undergoes SCSC transformation to form [Zn2(1,2-bdc)2(tpcb)0.5] (19) (tpcb
¼ tetrakis(40 -pridyl)-cyclobutane). Interestingly, the solvent is lost in this structural
conversion, and the interchain Zn  Zn distances changed to 3.57 and 4.43 Å [38].
Metal-Organic Frameworks for Photochemical Reactions 117

Fig. 12 The alignment of the photo-reactive olefin groups in 4PAH in Ag(I) 1D CPs

3.4 Stepwise Double [2þ2] Cycloaddition Reaction Assisted


by Ag  Ag Interactions

Biradha and coworkers employed Ag  Ag and Ag  π interactions to align the
diolefinic organic ligand, N,N0 -bis(3-(4-pyridyl)acryloyl)-hydrazine (4-PAH), and
the counter anion also showed its role to control over the selective orientation of the
olefinic groups for the photoreaction [39]. The 1D ladder CP, [Ag(4-PAH)(ClO4)]
(20), has parallel alignment of both the olefin groups, because of the Ag  Ag and
π  π interactions supported by the perchlorate anion. The crystal packing showed
the alignment of the olefin pairs in an asymmetric manner between the two linear
strands, assisted by the Ag  π (olefin) interactions as shown in Fig. 12. Upon UV
irradiation showed quantitative double [2þ2] photoreaction between the two linear
strands occurs by a stepwise mechanism as confirmed by the 1H-NMR studies.
Upon changing the counter anion to NO3, similar double-stranded pairs were
retained. But the interaction between the strands is so strong that the UV irradiation
furnished only a mono-cyclobutane product.

3.5 The Use of Conjugated Dienes for Double [2þ2]


Cycloaddition Reactions in 1D Polymers

Bis-pyridyl ligands with conjugated dienes are indeed interesting since they can
undergo double [2þ2] cycloaddition reactions. Once such ligand was the bis-pyridyl
118 R. Medishetty and J.J. Vittal

Fig. 13 The alignment of 1,3-pentadiene-3-one in Ag(I) 1D CPs

derivatives of 1,3-pentadiene-3-one used for making Ag(I) macromolecules and CPs


by Biradha and coworkers [39]. The reaction of this ligand with AgNO3 resulted in a
1D CP containing necklace-like loops made by Ag  Ag interactions as shown in
Fig. 13. The ligand shows cisoid–transoid conformation, but only the C═C bonds
with cisoid confirmation come closer between the interstrands. The other C═C
bonds are “out of phase” with each other. Interestingly, this solid exhibits quantita-
tive [2þ2] photoreaction, but the nature of the product is unknown. It is likely that
the product has 3D structure. However, changing the counter anion from NO3
to ClO4 showed completely different packing pattern due to cisoid–cisoid confor-
mation of the ligand. The resultant linear 1D strands are paired up by Ag  Ag
interactions to form ladder structure. This ladder compound is highly photo-reactive
to yield 100% double [2þ2] dimerized product as monitored by 1H-NMR spectros-
copy. Unfortunately, the structure of this 1D polymer was not confirmed [39].

3.6 Construction of Photo-Reactive 1D CP Using a Bioactive


Dicarboxylates

Usually, d10 metal ions are used to construct photo-reactive CPs. Here is an unusual
example of a transition metal ion to align a pair of olefin bonds in a biological active
molecule. A heterocyclic diolefin chelidonic acid (H2CDO) has been incorporated
into a photo-reactive 1D ladder polymer [Cu(CDO)(py)2(H2O)] (21) has been
constructed from CDO anions with Cu(II) in presence of pyridine [40]. The
neighboring six-membered pairs are aligned antiparallel with a distance of separa-
tion of 3.614 Å. Upon UV irradiation of this compound showed change in the color
from blue to green without losing the single crystalline nature, and the single crystal
Metal-Organic Frameworks for Photochemical Reactions 119

Fig. 14 Transformation of 21–22 by double photo-cycloaddition of a cyclic pairs

analysis of photo-irradiated crystal confirms the formation of photo-dimerized


product [Cu(EE1)(py)2(H2O)] (22) and resulted in the formation of a similar 1D
CP with cage-like ligand, EE1, and this ligand can be extracted and recrystallized as
sodium salt. The structural transformation is shown in Fig. 14.

4 1D to 2D Transformation

4.1 Synthesis of Unique Furan-Based Ligand

Usually, spacer ligands have been made photo-reactive by incorporating C═C bonds
in them. There are advantages as well as more challenges in making C═C containing
terminal ligands to be photo-reactive. A simple ligand that can be used for this
purpose is 4-styrylpyridine (4spy). The C═C double bonds can only be aligned in
head-to-tail fashion between different polymeric strands, and hence 4spy can be
aligned to undergo higher dimensional structural transformation using this versatile
ligand. Metal complex polymerization (0D to 1D) was described earlier in Sect. 2.2.
Here, an example for 1D to 2D structural transformation is described below.
120 R. Medishetty and J.J. Vittal

Fig. 15 Formation of 2D sheets from 1D CP by partial photochemical reaction

In the linear 1D CP [Cd(1,3-bdc)(4spy)2] (23) (1,3-H2bdc ¼ 1,3-benzenedi-


carboxylic acid, 4spy ¼ 4-styrylpyridine), the Cd(II) pairs are connected by four
1,3-bdc ligands as shown in Fig. 15 where one carboxylate bridges two Cd(II) centers
and the other carboxylate ligand chelates a Cd(II) center in a plane. The two axial
positions of the octahedral coordination environment of each Cd(II) atom are
filled by the 4spy ligands, and the pyridyl pairs are aligned parallel. However, the
olefinic bonds in the head-to-head pairs are misoriented and too far apart to
undergo a photochemical [2þ2] cycloaddition reaction. However, the 4spy
ligands from neighboring chains interdigitate in a head-to-tail fashion. A closer
look at the packing reveals that one of the two 4spy ligands is well aligned
with its partner from the neighboring strand with C  C distance of 3.86 Å.
In other words, 50% of the 4spy ligands can undergo HT dimerization above
and below the 1D strands [26]. Indeed, SCSC reaction occurs under UV light,
and the resultant compound [Cd(1,3-bdc)(4spy)(HT-ppcb)0.5] (24) (HT-ppcb ¼
1,3-bis(40-pyridyl)2,4-bis(phenyl)cyclobutane) has 2D sheet structure as shown
in Fig. 15 [41].
The ligand HT-ppcb was isolated and combined with CdCl2·2.5H2O,
1,3-H2bdc, and dilute H2O2 under hydrothermal conditions yielded another inter-
esting compound [Cd(H2O)(1,3-bdc)(bpbpf)]·H2O (25) (where bpbpf ¼ 2,4-bis
(4-pyridyl)-3,5-bis(phenyl)furan). During the hydrothermal reaction, the ligand
underwent in situ oxidation reaction in the presence of oxygen to form this furan
derivative [41].
Metal-Organic Frameworks for Photochemical Reactions 121

Fig. 16 Photo-reactivity of a triple-stranded 1D CP, 26

4.2 Photo-Reactivity of a Triple-Stranded CP

Apart from the usual ladder structures formed by a pair of polymeric strands
bridged by various anions as well as spacer ligands, one can also visualize
three polymer strands to assemble to form an unusual ladder structure. This
ladder can be considered as an intermediate in the self-assembly of 2D sheets.
Such triple-stranded ladders are of interests to investigate the photo-reactivity,
if C═C bonds in the three spacer ligands are aligned perfectly parallel within
Schmidt’s criteria.
Indeed, a photo-reactive triple-stranded ladder, [Pb3(bpe)3(O2C–CF3)4 (O2CCH3)2]
(26), containing parallel alignment of the three bpe spacer ligands has been
stumbled upon by Vittal and coworkers [42]. In this compound, Pb(II) in the center
strand is bridged by an acetate and a trifluoroacetate ligand, and the terminal Pb(II)
atoms of this 1D ribbon are chelated by trifluoroacetate anions, which prevent the
further assembly to form 2D CP. This compound showed interesting photo-
reactivity. UV irradiation of these single crystals within 10 h showed the expected
66% of dimerization of bpe due to the photoreaction between any two bpe
molecules across the strands and left the third one unreacted as shown in the time
versus percentage conversion plot as monitored by 1H-NMR studies in solution.
Surprisingly, further photo-irradiation showed quantitative photoreaction of bpe.
Moreover, grinding of the single crystals to powder form accelerates the second
step to completion in 40 h [42]. The authors suggested that there might be coopera-
tive movements of these triple strands during the process which aligned the
unreacted bpe between the intermolecular strands and thus the observed quantita-
tive [2þ2] cycloaddition as shown in Fig. 16. Such molecular movements have
been noted before [43, 44].
122 R. Medishetty and J.J. Vittal

Fig. 17 Schematic representation of transformation of 1D CP into 3D MOF through photo-


cycloaddition

5 1D to 3D Transformation in CPs

5.1 An Unusual Orientation of bpe Pairs for Photo-Reactivity

Tong and coworkers synthesized two photo-reactive 1D CPs, [M(1,2-chdc)(bpe)2


(H2O)2]·H2O (M ¼ Zn(II) (27), Mn(II) (28), 1,2-chdc ¼ trans-1,2-cyclohexanedi-
carboxylate) [45]. In these CPs, M(II) is present in octahedral coordination geometry,
which has been coordinated by two carboxylate O of chdc; two more sites are
coordinated by N of bpe and the coordination sphere is completed by water
molecules. The connectivity of the chdc ligands with two different metals leads
in the formation of 1D chain, and the bpe molecules are interestingly coordinated
from only one N, and the other side is H-bonded to the aqua ligands in the
neighboring chain. Similarly, the coordinated bpe in the next chain is, in turn,
also H-bonded to the first chain. Such complementary hydrogen bonds align the
double bonds of these two bpe in parallel with a C  C separation of 3.27 Å
as shown in Fig. 17. Upon UV irradiation, the C═C bonds from neighboring
bpe molecules form cyclobutane rings quantitatively without the loss of single
crystallinity. The new bonds formed in the direction normal to the propagation of
1D CP transformed product into 3D network structure with NbO topology [45].

6 Photo-Reactivity in 2D CPs

6.1 Photo-Reactivity of a Layer CP yielding an unusual


product

Here, we describe the photo-reactivity in an unusual crisscross alignment of the


reactive C═C bonds in adjacent CP layers [46]. In the 2D layer compound
[Cd2(fumarate)2(H2O)4] (29), the building block comprises a Cd(II) dimer. Each
Metal-Organic Frameworks for Photochemical Reactions 123

Fig. 18 Photo-reactivity of fumarates in 2D layered CP

Cd(II) is bonded to a chelating fumarate ion and two bridging fumarate anions
and forms a Cd2O2 node as shown in Fig. 18. Further connectivity of
Cd2(fumarate)4/2 dimer produced the expected (4,4) sheet structure. Each Cd(II)
is further bonded to two aqua ligands in a trans manner. These coordinated water
molecules are extensively hydrogen bonded to the adjacent layers. The staggered
arrangements of the adjacent layers in ABAB   fashion left no voids in the
structure. On the contrary, the olefinic groups of the fumarates in the adjacent
layers are separated by a distance of 3.38 Å in a crisscross fashion. Usually, such
arrangement will undergo pedal-like motion in the organic molecules and 1D CPs
but highly improbable in 2D layers. Hence, this cycloaddition reaction is to yield
rtct isomer. Interestingly, UV irradiation experiments yielded the expected unusual
rtct-cyclobutane-1,2,3,4-tetracarboxylate in the quantitative yield as monitored by
the appearance of the singlet at 3.31 ppm due to the cyclobutane C–H proton in the
1
H-NMR spectrum of the dissolved product. Due to ABAB   type of packing,
cyclobutane ring formation can yield either bilayer structure, if the reaction occurs
exclusively between two layers, or 3D structure if C–C bonds are formed randomly
on both sides of the layer. Since the crystals crashed during UV experiment, this
interesting information is not revealed [46].

6.2 2D to 2D0 Transformation

Miao and Zhu have synthesized a photo-reactive twofold interpenetrated


2D layered structure of [Cd2(CH3COO)2(3-sb)(bpe)2.5(H2O)]·4H2O (30) (3-sb ¼
3-sulfonatobenzoate) [47]. In the building block, two Cd(II) atoms are bridged
124 R. Medishetty and J.J. Vittal

Fig. 19 Photo-conversion of 31 to 32 upon UV irradiation

by two acetate anions separated by 3.85 Å. Each Cd(II) is coordinated to two
trans bpe ligands to form a 1D ladder chain. In one Cd(II), the hepta coordination
geometry is completed by a chelating 3-sb anion and an aqua ligand. Both the
acetate anions not only bridge the Cd(II) atoms but also chelate the second Cd(II)
atom. The equatorial position in the pentagonal bipyramid geometry of this Cd(II)
is completed by another bpe ligand bridging the neighboring ladder chains. This
leads to the formation of a brick wall type 2D structure, and the voids created
by this topology are minimized by twofold interpenetration. It is noted that the
[Cd2(OAc)2] unit dictates infinite parallel alignment of the bpe pairs and makes this
crystal photoactive. The 1H-NMR spectrum of the UV irradiated product showed
that all these aligned bpe molecules reacted. The product [Cd2(CH3COO)2(3-sb)
(bpe)0.5(4,40 -tpcb)2(H2O)]·4H2O (31) was independently synthesized by exposing
the reaction mixture of 1 in solution under sunlight for several hours and then
slowly evaporated at room temperature. The structure determination by X-ray
crystallography reveals the retention of interpenetrated 2D layer structures. Another
interesting aspect of this reaction is that 31 was obtained only with a high-power
mercury lamp (300 W, λ ¼ 367 nm) both in solution and the solid state after
irradiation of 8 h but not using a low power UV lamp (30 W) with λ ¼ 253.7 nm
for 9 days [47] (Fig. 19).

6.3 2D to 2D0 Transformation and Its Thermal


Cyclo-Reversion

The cleavage of the cyclobutane rings to the corresponding olefins is of interest


since reversible photochemical reaction can be investigated for potential appli-
cations in optical storage and sensing. However, such cleavage of cyclobutane rings
in CPs has never been observed. Recently, Vittal’s laboratory has succeeded in
cleaving the cyclobutane rings once formed in a 2D CP by thermal treatment [48].
Metal-Organic Frameworks for Photochemical Reactions 125

Fig. 20 Schematic representation of 2D to 2D photo-transformation of 33 and its thermal


reversibility

The following photo-reactive 2D CPs have been synthesized using an asymmetric


ligand, trans-2-(40 -pyridyl)vinylbenzoic acid (Hpvba): [Cd2(pvba)2(tbdc)(dmf)2]
(32), [Co2(pvba)2(tbdc) (dmf)2(H2O)2] (33), and [Ni2(pvba)2(tbdc)(dmf)2(H2O)2]
(34), where tpdc ¼ 2,3,5,6-tetra-bromobenzene dicarboxylate. In the 2D sheet,
[Cd2(O2C–C)4(pyridyl)2] core is the repeating unit connecting a pair of pvba
ligands which are aligned in head-to-tail parallel fashion in the orthogonal direction
and the tbdc ligands in the perpendicular direction. The C═C bonds in these pvba
ligand pairs are well aligned to undergo [2þ2] cycloaddition reaction quantitatively
in the solid state under UV light, yielding the cyclobutane derivative. This photo-
chemical reaction has been accompanied by a solid-state structural transformation
from one 2D structure to another as shown in Fig. 20 [48].
Further, the cyclobutane rings in the photodimerized product of 32 can be
cleaved by heating to 220 C for 12 h under vacuum in two different ways to
yield a mixture of trans and cis isomers of pvba quantitatively as monitored by
the 1H-NMR spectroscopy. This transformation as depicted in Fig. 20 appears to
be the first attempt to reverse the cyclized product in CPs. From the integration of
1
H-NMR spectrum, the ratio between the cis and trans isomers has been observed
to be 66.5% and 33.5%, respectively. Formation of other isomer of pvba has
also been supported by a new peak observed in photo luminescence at 370 nm,
which is different from trans pvba and the dimer. Under similar conditions, the
photoproducts of 33 and 34 showed scission of cyclobutanes up to 56% and
67%, respectively. Similar to 32, the new peaks also have been observed during
the scission of cyclobutane, which are related to cis isomer of pvba. However, the
intensities of these peaks are very small for integration from 1H-NMR spectral data.

6.4 Influence of 2D to 2D0 Structural Conversion


on the Gas Sorption Properties

Pore surface modification is the topic of current interest due to the ease in tuning
the properties for different applications. Kitagawa and coworkers have used [2þ2]
cycloaddition reaction as a post-synthetic method to modify the pore structure
126 R. Medishetty and J.J. Vittal

Fig. 21 The transformation of pores upon [2+2] photo-cycloaddition of bpe in 35

and surface area in a 2D sheet structure [49]. They synthesized a photo-reactive


twofold interpenetrating 2D CP, [Zn2(moip)2(bpe)2(DMF)2] (35) (H2moip ¼
5-methoxy-isophthalic acid). The 1D polymer formed by Zn2(moip)2 was further
linked by pairs of bpe ligands. These bpe ligands are oriented in parallel, and the
olefin groups are separated by 3.71 Å. In the resultant 3D structure obtained by
twofold interpenetration of these sheets, 1D channels are created with a cross
section of 3.2 Å  4.5 Å with the solvent-accessible space of 27.0% of the unit
cell volume. Upon UV irradiation of this compound, showed quantitative photo-
cycloaddition in an SCSC manner to form [Zn2(moip)2(rctt-tpcb)(DMF)2] (36).
During this process, the size of the 1D channel has been enlarged due to the
cyclobutane formation as shown in Fig. 21. Different gas sorption studies have
been performed, which showed no uptake of N2 at 77 K, which might be due to slow
diffusion rate. However, these compounds showed better adsorption of CO2 at
higher temperature, 195 K. The activated precursor sample showed three stepwise
adsorptions with the initial uptake up to 0.06 kPa and followed by a steep rise from
0.06 to 1 kPa and then gradual pore filling at further increase in pressure of CO2 and
also showed a hysteresis at the step 0.06 kPa. In contrast, the activated photoproduct
showed normal type I isotherms without any discontinuous steps. This observation
attributed the differences in the pore shape and the flexibility of the framework.
Hence, one could engineer the pore size and surface through photo-reactivity [49].

6.5 2D to 2D0 and 3D to 3D0 Structural Transformations

During the synthesis of 3D MOF, Natarajan’s laboratory separated an interesting


2D photo-reactive MOF, [Cd2(bpe)3(H2O)4(S2O3)2] (37), as a kinetic product [30].
In this solid-state structure, 1D chains have been formed between Cd(II) and
S (in μ2 mode) from thiosulfate, and the alternating Cd(II) atoms are coordinated
Metal-Organic Frameworks for Photochemical Reactions 127

Fig. 22 A diagram showing the photo-reactivity between the dangling bpe molecules in 37

by two types of bpe ligands. Among the first type is cross-linking the “Cd(II)-S”
chains, and the other type is coordinated to Cd(II) and hangs from other side
without coordinating. These “pendant” bpe from the adjacent “Cd(II)-S” chains
is perfectly aligned parallel, and the olefinic groups are separated by a distance of
4.02 Å. Upon UV irradiation of this compound showed complete photo-reactivity
of these aligned bpe molecules as shown in Fig. 22, which accounts for 67% of
whole bpe molecules in the compound. This reaction is expected to generate
another 2D layer structure through the formation of cyclobutane rings.
Similarly, thermodynamic product [Cd(bpe)S2O3] (38) was obtained by heating
the reaction solution of 38 at 60 C for 72 h. The structure is built from a tetranuclear
Cd4(S2O3)4 SBU made by Cd4S4 ring. The other two coordination sites in each
tetrahedral Cd(II) are occupied by bpe ligands. In other words, there are four
pairs of bpe ligands tetrahedrally oriented in this tetranuclear cluster to produce
fourfold interpenetrated diamondoid structure. Each bpe pair from two sulfur
bridged Cd(II) atoms is aligned in parallel but the olefinic groups are oriented in
antiparallel manner in two bpe pairs and separated by 3.79 Å. Upon UV irradiation
of this compound showed up to ~80% photo-reactivity and giving only rctt isomer
due to pedal motion of the C═C bonds [30].

6.6 2D to 3D Transformation

A unique 2D bilayered CP, [Mn2L2(H2O)2]·3H2O (39) (L ¼ E-5-(2-(pyridin-4-yl)


vinyl)isophthalic acid ligand), was reported by Wu and coworkers [50]. Each Mn(II)
atom with distorted octahedral coordination geometry is coordinated to a chelated
carboxylate O atoms, two syn and anti bridging O atoms from two different ligands,
one N atom from fourth ligand, and a water molecule. The Mn(II) carboxylates
form a 1D chain, which are further coordinated by pyridine groups from the
neighboring ligands to result in the formation a 2D bilayer. These bilayers arising
from the syn and anti bridging carboxylates are further packed in   AA   fashion
in 3D space. The olefinic groups of the ligands are aligned parallel within the
128 R. Medishetty and J.J. Vittal

Fig. 23 SCSC photo-transformation of a 2D CP, 39, to a 3D MOF

bilayer as well as between the bilayers in heat-to-tail manner. The distance between
the olefinic groups within the bilayers is 4.21 Å, but in contrast, the distance
between the olefinic groups between the layers is shorter (3.68 Å) as shown in
Fig. 23. By irradiation under UV light, this 2D bilayer structure showed quantitative
photo-cycloaddition between the layers in an SCSC manner and resulted in the
formation of a 3D MOF. Interestingly, the coordinated water molecule does not
destruct the single crystals during the cycloaddition reactions [50].

6.7 Transformation of 2D Interdigitated CP to 3D MOF

Terminal ligands have rarely been utilized to make photo-reactive CPs. In this
example, the use of 4spy in linking the 2D layers to form interpenetrated
3D structures by [2þ2] cycloaddition reaction is described [51]. Two isotypical
interdigitated photo-reactive CPs, [Zn2(cca)2(4spy)2] (40) and [Zn2(ndc)2(4spy)2]
(41), have been constructed from the paddle-wheel SBU, {Zn2(O2C–C)4}. The linear
dicarboxylate ligands 4-carboxycinnamate (cca) or 2,6-naphthalenedicarboxylate
(ndc) generate the 2D layer structure with (4,4) connectivity. The two apical positions
of each paddle-wheel SBU are coordinated by monodentate terminal 4spy ligands.
Due to long arms of this ligand, the 2D layers are interdigitated with each other in
the 3D space. The olefin groups of terminal 4spy ligands are aligned in heat-to-tail
fashion between the first and fourth layers with a distance of 3.85 Å. Of these
two 2D MOFs, only the cca derivate showed [2þ2] cycloaddition in an SCSC
fashion, and the final structure of photo reacted compound, [Zn2(cca)2(rctt-4-ppcb)]
(42) (where rctt-4-ppcb ¼ 1,3-bis(40 -pyridyl)-2,4-bis(phenyl)cyclobutane), has been
determined to be a triply interpenetrated MOF with α-Po topology as shown in
Fig. 24. But the single crystal was not reserved for 41 after the UV irradiation
experiment. It is demonstrated that the terminal photo-reactive ligands can also be
used for the structural transformation by [2þ2] photo-cycloaddition reactions.
Metal-Organic Frameworks for Photochemical Reactions 129

Fig. 24 Schematic representation of the transformation a 2D interdigitated CP to a triply


interpenetrated 3D MOF. The interpenetration is not shown

6.8 Post-Synthetic Modification to Fine-Tune the Sorption


Behavior of 2D Sheets

Kitagawa and coworkers have demonstrated how to control the reactivity and
pore surface of MOFs by photo-irradiation [52]. They have synthesized an inter-
digitated 2D sheet [Zn2(abdc)2(bpy)2(DMF)1.5] (43) using 5-amino-benzene-
1,3-dicarboxylate (abdc) and 4,40 -bipyridine (bpy). The amino group is exposed
on both sides of the pore which has been diazotized followed by azidation with
NaN3 to get azido derivative. Surprisingly, this MOF containing azide group is
thermally stable up to 150  C. But this is nonporous for oxygen sorption. On photo-
irradiation at 77 K, the adsorbed amount of oxygen was significantly improved
due to the loss of nitrogen and formation of triplet nitrene radicals in the pores.
Their results created a strategy to produce PCPs with reactive sites such as radicals
and carbenes which cannot be trapped by conventional synthetic conditions [52].
The details are depicted in Fig. 25.

7 Photo-Reactive 3D CPs

7.1 Post-Synthetic Modification of Double-Pillared MOFs


by [2þ2] Cycloaddition Reaction

Post-synthetic modification of double-pillared MOFs by [2þ2] photo-cycloaddition


was demonstrated in the following examples. Three photo-reactive double-pillared
MOFs, [Zn2(bpe)2(muco)2]·DMF·H2O (44), [Zn2(bpe)2(bdc)2]·DMF (45), and [Zn2(bpe)2
(fum)2]·H2O (46) (H2muco ¼ trans,trans-muconic acid, H2bdc ¼ 1,4-benzene
dicarboxylic acid, and H2fum ¼ fumaric acid), have been investigated [53]. In
these structures, the layer is formed by the {Zn2(O2C–C)4} SBU in which each
130 R. Medishetty and J.J. Vittal

Fig. 25 Schematic diagram illustrating the photoactivation of 2D CP, 43. (a) Photoactivation of
the azide moieties leads to the formation of triplet nitrenes on the pore surface. (b) Photochemical
trapping of physisorbed oxygen molecules in the framework of dried 43. This figure is reproduced
from [52] with permission

Zn(II) is chelated by a carboxylate and bridged by two carboxylates. These


2D layers are further double-pillared by the bpe ligands at the axial positions of
Zn(II) atoms which are separated by 3.96 Å, to produce the 3D structure. This
arrangement assists the bpe pairs to orient in parallel with the olefin groups
separated by 3.79 Å (Fig. 26). Upon irradiation under UV light, all these MOFs
showed quantitative cycloaddition reaction across the double pillars as confirmed
by 1H-NMR spectral studies. Among the three, MOFs containing muco and bdc
ligands (44 and 45) underwent SCSC structural transformation. It has been shown
that it is possible to modify the backbone of the spacer ligand without affecting the
topology or porosity of the MOF during this structural transformation [53].
Metal-Organic Frameworks for Photochemical Reactions 131

Fig. 26 Schematic representation of post-synthetic modification of 3D MOF using [2þ2] photo-


dimerization reaction

Fig. 27 A portion of the layer formed by Cd2(fumarate)2 (left) and the photo-reactive double-
pillared MOF (47) (right)

A Cd(II) fumarate compound [Cd2(bpe)2(fum)2] (47) similar to the Zn(II)


analog with double-pillared MOF structure shown in Fig. 27 was also shown to
be photo-reactive [54].

7.2 Photo-Reactive Double-Pillared MOFs Containing


Conjugated Olefin Bonds

Conjugated C═C bonds containing spacer ligands are interesting due to many
possibilities these ligands can offer in terms of photo-reactivity. When both
C═C bonds are aligned “face-to-face” in parallel, the ligand can undergo double
[2þ2] cycloaddition reaction. One C═C bond from each ligand can also take
132 R. Medishetty and J.J. Vittal

Fig. 28 “Out of phase” photo-reactivity of bpeb in 48.

part in the cyclobutane formation, if the ligand pair is arranged in “out-of-phase”


manner. Both situations have been encountered by Lang and coworkers by
using the conjugated ligand, 1,4-bis[2-(4-pyridyl)ethenyl]benzene (bpeb) [55].
Two photo-reactive MOFs [Zn4(μ3-OH)2(5-sipa)2(1,4-bpeb)2]·4H2O (48) and
[Cd2(1,3-pda)2(1,4-bpeb)2] (49) (5-H3sipa ¼ 5-sulfoisophthalic acid and 1,3-
H2pda ¼ 1,3-phenylenediacetic acid) have been synthesized.
In the case of Zn(II) MOF, a 2D layer [Zn4(μ3-OH)2(5-sipa)2] with (6,3)
connectivity related CdI2 is formed by linking six equivalent 5-sipa with three
equivalent {Zn4(μ3-OH)2}. The axial positions of these {Zn4(μ3–OH)2} sheets are
further pillared by two pairs of bpeb ligands to form the double-pillared 3D network
structure. However, each bpeb pillar is arranged in between two other bpeb ligands,
where one C═C bond is aligned “in-phase” and the other in “out-of-phase,” and
all the olefin groups satisfy the topochemical distance criteria. However, the
“in-phase” C═C bond pairs are crisscrossed and the “out-of-phase” double bonds
have parallel orientation. As a result, UV irradiation of this compound showed
quantitative photoreaction between the “out of phase” C═C pairs only and resulted
in the formation of another 3D structure in an SCSC fashion (Fig. 28).
Similarly in the (4,4) sheets formed by the {Cd2(1,3-pda)4}, the bpeb pairs
bonded to the Cd(II) atoms are extended from either side of the sheets to link the
neighboring {Cd2(1,3-pda)4} sheets to form another double-pillared 3D structure.
Metal-Organic Frameworks for Photochemical Reactions 133

Fig. 29 “In-phase” photo-reactivity of bpeb in 49

The bpeb ligand pairs are aligned in “in-phase” manner with the distances between
C═C bonds, 3.886 and 4.023 Å. The structural transformation occurs in an SCSC
manner upon UV irradiation of these crystals with the quantitative formation of
double cyclobutane rings across bpeb pairs (Fig. 29). This is another elegant
example for the post-synthetic modification of double-pillar ligands into a single
pillar by [2þ2] cycloaddition reaction [55].

7.3 Host–Guest Reactions in Pillared-Layer MOFs

An interesting alignment between the coordinated and guest bpe molecules in the
channels that leads to photo-reactivity was described by Gao’s laboratory [56].
A photo-reactive pillared-layer CP [Mn2(HCO2)3(bpe)2(H2O)2] ClO4·H2O·bpe (50)
has the layer structure formed by Mn2(HCO2)3 linked by bpe pillars. Each Mn(II)
is linked to three others through bridging formate ligands and an aqua ligand
to form a layer, and the two axial sites are occupied by the bpe ligands. Two
different types of channels formed in this 3D framework structure are occupied
134 R. Medishetty and J.J. Vittal

Fig. 30 The alignment of bpe in the host–guest fashion in 50

by the anions, lattice water and guest bpe molecule. One of the coordinated
bpe ligands is aligned parallel with a guest bpe (Fig. 30) stabilized by the
O–H  N hydrogen bonds with the aqua ligands. This compound is expected to
yield 66.7% cyclobutane product based on this host–guest interaction, and indeed
about 60% photo-reactivity was observed under UV light as monitored by the
1
H-NMR spectroscopy [56]. This appears to be the first [2þ2] solid-state photo-
reaction of a guest molecule with the framework.
Kitagawa and coworkers have recently described an interesting but different
host–guest-polymerization concept to synthesize cross-linked polymers such as
polystyrene, methylmethacrylate, and vinylacetate with pseudo-crystallinity in
non-photochemical route [57]. In order to achieve this, they have first incorporated
the cross-linker 2,5-divinyl-benzene-1,4-dicarboxylate (DVTP) into the porous CP
[Cu(DVTP)(triethylenediamine)0.5] (51). The host framework containing porous
channels with dangling vinyl groups provides a suitable environment for radical
polymerization of these monomers as shown in Fig. 31. This is obviously different
from photopolymerization by [2þ2] cycloaddition reaction.

7.4 Cis–Trans Isomerization on the Gas Adsorption


Property of MOFs

Cis–trans isomerization of azobenzene has widely been exploited in photo-


responsive materials [58, 59]. The pore structure of the MOF [Zn2(bdc)2(triethyle-
nediamine)] (52) (pore size, 7.5 Å  7.5 Å) has been shown to be deformed by
the inclusion of certain aromatic guest molecules. This structural dynamic property
of the MOF and the photo behavior of azobenzene have been nicely combined
Metal-Organic Frameworks for Photochemical Reactions 135

Fig. 31 Radical polymerization on the MOF surfaces. (a) Molecular structure of the
DVTP ligand. (b) Monomers used in the host–guest cross-polymerizations. (c) Cross-linkable
MOF [Cu(DVTP)x(terephthalate)1x(triethylenediamine)0.5] (1x shown in light blue) prepared
in solid solution fashion. (d) Host 1x accommodates vinyl monomers in its nanocavities.
(e) Heat triggers radical cross-polymerization in the presence of AlBN initiator, yielding 1x
polymer nanocomposites. (f) Selective decomposition of the MOF matrix generates highly ordered
cross-linked polymers. This figure is reproduced from [57] with permission

Fig. 32 Reversible transformation of 52 by photo isomerization of azobenzene

to study the cis–trans isomerization of azobenzene inside the pores of 52. The
change in the structure of MOF due to UV light triggered reversible trans to
cis form and thermal isomerization of cis to trans form resulting in a drastic
switching of the adsorption property of the MOF. The isomerization under UV
light can easily be followed by the disappearance of the frequency at 690 cm1 due
to the trans isomer and appearance of a new frequency at 697 cm1 due to the
formation of cis isomer in the IR spectroscopy. This isomerization is accompanied
by the changes of the framework structure from orthorhombic to the tetragonal
form (Fig. 32). This dramatic change in the structure also influences the nitrogen
adsorption behavior [60].
136 R. Medishetty and J.J. Vittal

Fig. 33 Partial formation of ladderane in 53 upon UV irradiation. The disordered atoms have
been removed for clarity

7.5 Photo-Reactive Lanthanide-MOFs (Ln-MOF)

Lanthanum(III) atoms have rarely been used to study the photo-reactivity in CPs.
Michaelides and coworkers explored the photo-reactivity of muconate ligands
bonded to Er(III) and Y(III) atoms. Muconate is another linear spacer ligand with
two conjugated C═C bonds in the backbone which presents different possibilities
of photoproducts, including mono [2þ2] cycloaddition or [4þ4] cycloaddition
products. The transformation of cyclobutane to cyclooctadiene by Cope rearrange-
ment is another possibility. Further, it may also result in the formation of a
highly strained ladderane structure by double [2þ2] cycloaddition reaction.
The two Ln(III) MOFs [Ln2(muco)3(H2O)6]·2.5H2O (H2muco ¼ trans,trans-
muconic acid), Ln(III) ¼ Er(III) (53) and Y(III) (54) have the face-to-face parallel
alignment between the muconate ligands templated by the Ln(III) atoms in which
the olefins are separated by 3.60 Å [61]. On UV irradiation, these compounds
showed photo-reactivity. As there are many possibilities in the photoreaction
of muconate ligand as discussed before, identifying the photoproduct is one of
the challenging tasks in the absence of SCSC conversion. Usually, 1H-NMR
spectroscopy has been used to monitor the nature of products in solution. Due to
the insoluble nature of these products, they are normally digested with acids
to solubilize the organic moiety, and there is a possibility for these products
to rearrange from one structure to another under these conditions in solution.
Interestingly, the photo-irradiation of Er(III) showed formation of ladderane
structure with double [2þ2] photo-cycloaddition reaction as shown in Fig. 33,
which has been confirmed by partial SCSC transformation, where the single
crystals are stable up to 55% of photoreaction. This highly strained ladderane
formation has also been supported by other experimental evidences. This may
be due to the generation of the stress created during the photoreaction, the single
crystals were collapsed after 55% photoreaction [61]. Similarly, isomorphs of
the above structures have also been investigated by the same authors [62]. In all
these compounds, the separation distance between the olefin bonds varies from
3.7 to 4.4 Å and also not aligned in parallel. However, all these compounds
participated in the photoreaction as confirmed by 1H-NMR spectral studies.
Metal-Organic Frameworks for Photochemical Reactions 137

8 [2þ2] Cycloaddition Reaction to Monitor the


Structural Transformations in CPs

8.1 Anisotropic Movements of 1D CPs by Desolvation

Since close proximity is essential for the [2þ2] cycloaddition reaction, only ladder
structure in which the C═C bond incorporated spacer ligands occupy the poles can
be photo-reactive among the 1D CPs. Hence, the existence or the formation of
ladder structures with these photo-reactive ligands can easily be monitored by
the 1H-NMR spectroscopy for the formation of cyclobutane rings after irradiating
them under UV light, if the single crystal data are not available. The photo-
reactivity of the ladder CPs has been used as a tool to understand the anisotropic
molecular movements of CPs [43]. A 1D CP, [Ag(μ-bpe)-(H2O)](TFA)·CH3CN
(55), has been trapped as a kinetic product during the synthesis of 1D ladder CP,
[Ag2(bpe)2(TFA)2] (56). In the linear CP, the Ag(I) has a T-shaped geometry
from two bpe spacers and an aqua ligand. The hydrogen atoms of the coordinated
water molecule are H-bonded to the O atoms of noncoordinated TFA anions
and form a 2D H-bonded sheet-like structure. The olefin groups of the bpe are
far away (5.15 Å) to satisfy the topochemical distance criterion. Surprisingly,
the aqua ligand can easily be lost in air along with CH3CN, and the single crystals
were destroyed very easily. However, UV irradiation of this desolvated powder
showed quantitative formation of cyclobutane rings as confirmed by 1H-NMR data.
The loss of coordinatively bonded aqua ligands at such a low temperature suggests
some kinds of structural transformation taken place in the solid state. This photo-
reaction confirms the formation of a ladder-like structure from the linear chains. All
these suggest that the linear chain is a kinetic product and cooperative anisotropic
molecular movements of the adjacent 1D chains occur during the desolvation
process leading to the formation of the thermodynamically stable ladder CP
(Fig. 34), which is stabilized by Ag  Ag and face-to-face π  π interactions.

8.2 Transformation of a Linear CP to a Ladder Structure


by Thermal Dehydration

Another structural rearrangement of a linear CP to a ladder structure was also


successfully monitored by the [2þ2] cycloaddition reactions [44]. The Cd(II)
center in the linear CP [Cd(bpe)(CH3COO)2(H2O)] (57) has hepta-coordinated
pentagonal bipyramidal geometry coordinated to two bpe ligands in trans fashion
and chelated by two acetate groups and a water molecule. These linear chains are
arranged in parallel one below another to form a plane with the help of H-bonds
between the coordinated water molecule and chelated acetate O atoms from the
138 R. Medishetty and J.J. Vittal

Fig. 34 Rearrangement of a linear CP to a photo-reactive ladder structure by desolvation

Fig. 35 Structural transformation of 57 upon desolvation and its photo-reactivity upon UV irradiation

adjacent chains as shown in Fig. 35. A close examination of the packing reveals
that the closest bpe pairs are crisscrossed and the distance between these olefinic
groups is 4.33 Å and hence expected to be a photo stable solid [63]. However,
this compound showed partial photo-cycloaddition of bpe (33%) when exposed
Metal-Organic Frameworks for Photochemical Reactions 139

to UV irradiation for 30 h, and longer irradiation did not increase the formation
of cyclobutane rings as monitored by 1H-NMR spectroscopy. Interestingly, the
TGA showed that the loss of coordinated water occurs in the range, 67–106 C.
The loss of coordinated water can bring the adjacent chains closer to form a
new bond between Cd(II) and oxygen atom of the adjacent chains, which are
separated by 4.44 Å. During this structural transformation, formation of a stable
structure might have forced the loss of coordinated water below its boiling point
as observed by TGA. The dehydrated product exhibited 100% photo-reactivity
when exposed to UV light 30 h. This photo-cycloaddition strongly supports
the migration of acetates after the loss of coordinated solvent, in bringing the
adjacent chains closer to form a ladder structure over 2D sheets. Therefore, the
formation of cyclobutane rings can be used as a tool to follow the alignment of
the double bonds during structural transformation.

8.3 Loss of Water Chains Transforms a Linear CP


to a Ladder Structure

During the synthesis of [Pb2(μ-bpe)2(μ-O2C–C6H5)2(O2C–C6H5)2] (13) (see Sect. 3.1),


a linear CP [Pb(μ-bpe)(O2C–C6H5)2]·2H2O (58) was trapped as a kinetic product [34].
In the solid-state structure, the guest water molecules aggregate to form a
hydrogen-bonded 1D polymer. When this water aggregate was removed by thermal
dehydration, it transforms to the photo-reactive ladder structure as confirmed
by PXRD and UV irradiation experiments (Fig. 36). This work illustrates the
generality of the design principles of synthesizing double-stranded ladder-like 1D
CPs to align infinite C═C bond pairs for solid-state photo-reactivity by clipping
hemidirected Pb(II) centers in a serial fashion.

9 Photopolymerization of Alkenoate Systems in CPs

Solid-state polymerization of Li(I), Na(I), and K(I) acrylates has been studied for
many years, and such reactions have been induced by 60Co γ-rays. Later, the
polymerization of Ca(II), Ba(II), and Zn(II) methacrylates, which are of 1D
and 2D CPs, was also reported [64–72]. In addition, some of these compounds
also undergo polymerization by heating. However, the final structures of these
compounds are not known since they do not retain their single crystallinity after
the polymerization reaction.

10 Photo-Reactivity in Solution

A number of reactions conducted in solution including hydrothermal/solvothermal


process produced CPs and MOFs containing photoproducts as evidence for the
photo-reactivity in solution. It is clear that the metal ions, anions, and the co-ligands
140 R. Medishetty and J.J. Vittal

Fig. 36 Transformation of a linear CP, 17, to a photo-reactive ladder structure, 13, upon dehydration

assist in the formation of the stereoselective formation of cyclobutane derivatives


in the novel solid-state structures. Once formed, the organic photoproduct can
combine with the metal ions, anions, and co-ligands to form interesting CPs and
MOFs [49–52]. But this area of research is out of scope of this chapter and hence
not discussed further.

11 Conclusion

In this chapter, various photochemical reactions involving CPs and MOFs


have been described. These include structural transformations, post-synthetic
modification of surfaces to create reactive sites, radicals which are not possible
by convention synthetic routes, polymerization on the surfaces, and cis–trans
isomerization of the guest molecules in the cavities using the dynamic behavior
of the structures. Further, [2þ2] cycloaddition reaction has been used as a tool
Metal-Organic Frameworks for Photochemical Reactions 141

to monitor the changes in the solid-state structures in the absence of crystal


structures. A number of polymerization reactions such as alkenoates have been
investigated. The structures of the resultant products were hampered by the lack
of suitable structural tool other than single crystal X-ray crystallography. On the
other hand, most of the examples discussed in this chapter do maintain the
single crystals at the end of the reactions making them amenable for structural
elucidation to understand the reactivity completely.

Acknowledgements We thank the National University of Singapore for their continuous support.
The Ministry of Education, Singapore, is gratefully thanked for financial support through NUS
FRC Grant R-143-000-439-112. We also thank the Nature Publications for providing the copyright
permissions for Figs. 25 and 31.

References

1. Ramamurthy V, Venkatesan K (1987) Photochemical reactions of organic crystals. Chem Rev


87(2):433–481
2. Theocharis CR, Jones W (1987) Organic solid state chemistry. In: Desiraju GR (ed) Organic
solid state chemistry. Elsevier, Amsterdam, pp 47–68
3. Green BS, Lahav M, Rabinovich D (1979) Asymmetric synthesis via reactions in chiral
crystals. Acc Chem Res 12(6):191–197
4. Jones W (ed) (1997) Organic molecular solids: properties and applications. CRC, Boca Raton
5. Toda F (ed) (2005) Organic solid state reactions. Topics in current chemistry, vol 254.
Springer, Berlin
6. Keating AE, Garcia-Garibay MA (1998) Organic and inorganic photochemistry. In:
Ramamurthy V, Schanze KS (eds) Organic and inorganic photochemistry. Marcel Dekker,
New York, pp 195–248
7. Tanaka K, Toda F (2000) Solvent-free organic synthesis. Chem Rev 100(3):1025–1074
8. Toda F (1995) Solid state organic chemistry: Efficient reactions, remarkable yields, and
stereoselectivity. Acc Chem Res 28(12):480–486
9. Lee-Ruff E, Mladenova G (2003) Enantiomerically pure cyclobutane derivatives and their use
in organic synthesis. Chem Rev 103(4):1449–1484
10. Namyslo JC, Kaufmann DE (2003) The application of cyclobutane derivatives in organic
synthesis. Chem Rev 103(4):1485–1538
11. Nishimura J, Nakamura Y, Hayashida Y, Kudo T (2000) Stereocontrol in cyclophane synthesis:
a photochemical method to overlap aromatic rings. Acc Chem Res 33(10):679–686
12. Cohen MD, Schmidt GMJ, Sonntag FI (1964) Topochemistry. Part II. The photochemistry of
trans-cinnamic acids. J Chem Soc 384:2000–2013
13. Schimdt GMJ (1971) Photodimerization in the solid state. Pure Appl Chem 27:647–678
14. Desiraju GR, Vittal JJ, Ramanan A (eds) (2011) Crystal engineering. World Scientific,
Singapore
15. Kole GK, Vittal JJ (2013) Solid-state reactivity and structural transformations involving
coordination polymers. Chem Soc Rev 42(4):1755–1775
16. MacGillivray LR, Papaefstathiou GS, Friščić T, Hamilton TD, Bučar D-K, Chu Q, Varshney
DB, Georgiev IG (2008) Supramolecular control of reactivity in the solid state: from templates
to ladderanes to metal – organic frameworks. Acc Chem Res 41(2):280–291
17. Biradha K, Santra R (2013) Crystal engineering of topochemical solid state reactions. Chem
Soc Rev 42(3):950–967
142 R. Medishetty and J.J. Vittal

18. Nagarathinam M, Peedikakkal AMP, Vittal JJ (2008) Stacking of double bonds for photo-
chemical [2þ2] cycloaddition reactions in the solid state. Chem Commun 42:5277–5288
19. Nagarathinam M, Vittal JJ (2006) A rational approach to crosslinking of coordination
polymers using the photochemical [2þ2] cycloaddition reaction. Macromol Rapid Commun
27(14):1091–1099
20. Vittal JJ (2007) Supramolecular structural transformations involving coordination polymers
in the solid state. Coord Chem Rev 251(13–14):1781–1795 (special issue)
21. Praetorius P (1909) Thesis, Halle
22. Praetorius P, Korn F (1910) Ber Dtsch Chem Ges 43:2744
23. Stobbe H, Färber E (1925) Ber Dtsch Chem Ges 58:1548
24. Alcock NW, de Meester P, Kemp TJ (1979) Solid-state photochemistry. Part 1. Nature of the
stereocontrol in the photodimerisation of dibenzylideneacetone by UO22þ ion: Crystal and
molecular structure of trans-dichlorobis(trans, trans-dibenzylideneacetone)dioxouranium(VI)
and of its acetic acid solvate. J Chem Soc Perkin Trans 2(7):921–926
25. Theocharis CR (1987) Co-ordination polymers based on 2,5-dibenzylidenecyclopentanone,
which are photochemically cross-linkable. J Chem Soc Chem Commun 0(2):80–81
26. Chu Q, Swenson DC, MacGillivray LR (2005) A single-crystal-to-single-crystal transformation
mediated by argentophilic forces converts a finite metal complex into an infinite coordination
network. Angew Chem Int Ed 44(23):3569–3572
27. Harada J, Ogawa K (2001) Invisible but common motion in organic crystals: A pedal motion in
stilbenes and azobenzenes. J Am Chem Soc 123(44):10884–10888
28. Dutta S, Bucar D-K, Elacqua E, MacGillivray LR (2013) Single-crystal-to-single-crystal
direct cross-linking and photopolymerisation of a discrete Ag(I) complex to give a 1D poly-
cyclobutane coordination polymer. Chem Commun 49(11):1064–1066
29. Peedikakkal AMP, Koh LL, Vittal JJ (2008) Photodimerization of a 1D hydrogen-bonded
zwitter-ionic lead(ii) complex and its isomerization in solution. Chem Commun 4:441–443
30. Paul AK, Karthik R, Natarajan S (2011) Synthesis, structure, photochemical [2þ2] cycloaddi-
tion, transformation, and photocatalytic studies in a family of inorganic–organic hybrid cadmium
thiosulfate compounds. Cryst Growth Des 11(12):5741–5749
31. Peedikakkal AMP, Vittal JJ (2008) Solid-state photochemical [2þ2] cycloaddition in a
hydrogen-bonded metal complex containing several parallel and crisscross C═C bonds.
Chem Eur J 14(17):5329–5334
32. Toh NL, Nagarathinam M, Vittal JJ (2005) Topochemical photodimerization in the coordina-
tion polymer [{(CF3CO2)(μ-O2CCH3)Zn}2(μ-bpe)2]n through single-crystal to single-crystal
transformation. Angew Chem Int Ed 44(15):2237–2241
33. Nagarathinam M, Vittal JJ (2010) Solid-state synthesis of coordination polymers for [2þ2]
photoreactions by grinding. Aust J Chem 63(4):589–595
34. Kole GK, Peedikakkal AMP, Toh BMF, Vittal JJ (2013) Solid-state structural transformations
and photoreactivity of 1D-ladder coordination polymers of PbII. Chem Eur J 19(12):
3962–3968
35. Liu D, Li N-Y, Lang J-P (2011) Single-crystal to single-crystal transformation of 1D coordi-
nation polymer via photochemical [2þ2] cycloaddition reaction. Dalton Trans 40(10):
2170–2172
36. Papaefstathiou GS, Zhong Z, Geng L, MacGillivray LR (2004) Coordination-driven self-
assembly directs a single-crystal-to-single-crystal transformation that exhibits photocontrolled
fluorescence. J Am Chem Soc 126(30):9158–9159
37. Papaefstathiou GS, Georgiev IG, Friscic T, MacGillivray LR (2005) Directed assembly and
reactivity of olefins within a one-dimensional ladder-like coordination polymer based on a
dinuclear Zn(ii) platform. Chem Commun 31:3974–3976
38. Ou Y-C, Liu W-T, Li J-Y, Zhang G-G, Wang J, Tong M-L (2011) Solvochromic and
photodimerization behaviour of 1D coordination polymer via single-crystal-to-single-crystal
transformation. Chem Commun 47(33):9384–9386
Metal-Organic Frameworks for Photochemical Reactions 143

39. Santra R, Banerjee K, Biradha K (2011) Weak AgAg and Ag[small pi] interactions in
templating regioselective single and double [2þ2] reactions of N, N[prime or minute]-bis
(3-(4-pyridyl)acryloyl)-hydrazine: Synthesis of an unprecedented tricyclohexadecane ring
system. Chem Commun 47(38):10740–10742
40. Eubank JF, Kravtsov VC, Eddaoudi M (2007) Synthesis of organic photodimeric
cage molecules based on cycloaddition via metal–ligand directed assembly. J Am Chem Soc
129(18):5820–5821
41. Liu D, Lang J-P, Abrahams BF (2013) Stepwise ligand transformations through [2þ2]
photodimerization and hydrothermal in situ oxidation reactions. Chem Commun 49(26):
2682–2684
42. Peedikakkal AMP, Vittal JJ (2010) Solid-state photochemical behavior of a triple-stranded
ladder coordination polymer. Inorg Chem 49(1):10–12
43. Nagarathinam M, Vittal JJ (2006) Anisotropic movements of coordination polymers upon
desolvation: solid-state transformation of a linear 1D coordination polymer to a ladderlike
structure. Angew Chem Int Ed 45(26):4337–4341
44. Nagarathinam M, Vittal JJ (2008) Photochemical [2þ2] cycloaddition as a tool to study a
solid-state structural transformation. Chem Commun 4:438–440
45. Ou YC, Zhi DS, Liu WT, Ni ZP, Tong ML (2012) Single‐crystal‐to‐single‐crystal transforma-
tion from 1D staggered‐sculls chains to 3D NbO‐type metal‐organic framework through [2þ2]
photodimerization. Chem Eur J 18(24):7357–7361
46. Michaelides A, Skoulika S, Siskos MG (2004) Assembly of a photoreactive coordination
polymer containing rectangular grids. Chem Commun 21:2418–2419
47. Miao X-H, Zhu L-G (2010) Regiocontrolled [2þ2] photodimerization of Cd(II) metal
complexes in both solution and solid state. Dalton Trans 39(6):1457–1459
48. Chanthapally A, Kole GK, Qian K, Tan GK, Gao S, Vittal JJ (2012) Thermal cleavage of
cyclobutane rings in photodimerized coordination-polymeric sheets. Chem Eur J 18(25):
7869–7877
49. Sato H, Matsuda R, Mir MH, Kitagawa S (2012) Photochemical cycloaddition on the
pore surface of a porous coordination polymer impacts the sorption behavior. Chem Commun
48(64):7919–7921
50. Xie M-H, Yang X-L, Wu C-D (2011) From 2D to 3D: a single-crystal-to-single-crystal
photochemical framework transformation and phenylmethanol oxidation catalytic activity.
Chem Eur J 17(41):11424–11427
51. Medishetty R, Koh LL, Kole GK, Vittal JJ (2011) Solid-state structural transformations from
2D interdigitated layers to 3D interpenetrated structures. Angew Chem Int Ed 50(46):
10949–10952
52. Sato H, Matsuda R, Sugimoto K, Takata M, Kitagawa S (2010) Photoactivation of a
nanoporous crystal for on-demand guest trapping and conversion. Nat Mater 9(8):661–666
53. Mir MH, Koh LL, Tan GK, Vittal JJ (2010) Single-crystal to single-crystal photochemical
structural transformations of interpenetrated 3D coordination polymers by [2þ2] cycloaddi-
tion Reactions13. Angew Chem Int Ed 49(2):390–393
54. Michaelides A, Skoulika S, Siskos MG (2008) Designed self-assembly of a reactive metal-
organic framework with quasi [small alpha]-Po topology. CrystEngComm 10(7):817–820
55. Liu D, Ren Z-G, Li H-X, Lang J-P, Li N-Y, Abrahams BF (2010) Single-crystal-to-single-
crystal transformations of two three-dimensional coordination polymers through regio-
selective [2þ2] photodimerization reactions. Angew Chem Int Ed 49(28):4767–4770
56. Wang X-Y, Wang Z-M, Gao S (2007) A pillared layer MOF with anion-tunable magnetic
properties and photochemical [2þ2] cycloaddition. Chem Commun 11:1127–1129
57. Distefano G, Suzuki H, Tsujimoto M, Isoda S, Bracco S, Comotti A, Sozzani P, Uemura T,
Kitagawa S (2013) Highly ordered alignment of a vinyl polymer by host–guest cross-
polymerization. Nat Chem 5(4):335–341
58. Beharry AA, Woolley GA (2011) Azobenzene photoswitches for biomolecules. Chem Soc
Rev 40(8):4422–4437
144 R. Medishetty and J.J. Vittal

59. Kumar GS, Neckers DC (1989) Photochemistry of azobenzene-containing polymers. Chem


Rev 89(8):1915–1925
60. Yanai N, Uemura T, Inoue M, Matsuda R, Fukushima T, Tsujimoto M, Isoda S, Kitagawa S
(2012) Guest-to-host transmission of structural changes for stimuli-responsive adsorption
property. J Am Chem Soc 134(10):4501–4504
61. Michaelides A, Skoulika S, Siskos MG (2011) Photoreactive 3D microporous lanthanide
MOFs: formation of a strained ladderane in a partial single crystal-to-single crystal manner.
Chem Commun 47(25):7140–7142
62. Michaelides A, Skoulika S, Siskos MG (2013) 2D And 3D photoreactive lanthanide MOFs of
trans, trans-muconic acid. Chem Commun 49(10):1008–1010
63. Schmidt GMJ (1971) Pure Appl Chem 27:647–678
64. Vela MJ, Snider BB, Foxman BM (1998) Solid-state polymerization of aquabis(3-butenoato)
calcium. Chem Mater 10(10):3167–3171
65. Inoki M, Akutsu F, Kitayama Y, Kasashima Y, Naruchi K (1998) Thermal polymerizations
of alkali and alkaline earth 4-vinylbenzoates in bulk. Macromol Chem Phys 199(4):619–623
66. Kudoh M, Akutsu F, Nakaishi E, Kobayashi T, Naruchi K, Miura M (1994) Thermal polymer-
ization of alkali and zinc methacrylates in the solid state. Macromol Rapid Commun 15(3):
239–242
67. Bowden MJ, O’Donnell JH, Sothman RD (1969) Radiation induced solid state polymerization
of derivatives of methacrylic acid. VI. Polymerization of barium methacrylate dihydrate
during irradiation. Macromol Chem Phys 122(1):186–195
68. O’Donnell JH, Sothman RD (1968) Radiation-induced, solid-state polymerization of deriva-
tives of methacrylic acid. I. Postirradiation polymerization of zinc methacrylate. J Polym Sci
Part A Polym Chem 6(5):1073–1086
69. Costaschuk FM, Gilson DFR, St. Pierre LE (1971) The solid-state polymerization of hydrated
barium methacrylate. Macromolecules 4(1):16–19
70. Costaschuk FM, Gilson DFR, St. Pierre LE (1970) Solid-state polymerization of hydrated
calcium acrylate. Macromolecules 3(4):393–397
71. Alcock NW, de Meester P, Kemp TJ (1979) Solid-state photochemistry. Part 1. Nature of
the stereocontrol in the photodimerisation of dibenzylideneacetone by UO22þ ion: Crystal
and molecular structure of trans-dichlorobis(trans, trans-dibenzylideneacetone)dioxouranium
(VI) and of its acetic acid solvate. J Chem Soc Perkin Trans 2 0(7):921–926
72. Vela MJ, Buchholz V, Enkelmann V, Snider BB, Foxman BM (2000) Solid-state polymeriza-
tion of bis(but-3-enoato)zinc: The generation of a stereoregular oligomer. Chem Commun
0(22):2225–2226
Struct Bond (2014) 157: 145–166
DOI: 10.1007/430_2013_107
# Springer-Verlag Berlin Heidelberg 2013
Published online: 29 May 2013

Metal–Organic Frameworks for


Second-Order Nonlinear Optics

Shaowu Du and Huabin Zhang

Abstract In this chapter, we highlight recent advances and perspectives in the


design and synthesis of NLO-active metal–organic frameworks (MOFs). Some of
these compounds show impressive second harmonic generation (SHG) responses
and may have potential use in NLO-applications. Noncentrosymmetric MOFs can
be synthesized mainly by using chiral ligands, unsymmetrical achiral ligands, and
mixed metal ions. Organic ligands containing donor-π-acceptor systems can be
incorporated into MOFs to improve the SHG activity of the bulk materials through
push–pull effect. Some diamondoid or octupolar MOFs have been shown to exhibit
excellent SHG properties. Besides, a number of MOFs with 2D grid-type or 1D
helical chain structures display good SHG activity.

Keywords Diamondoid net  Metal–organic framework  Noncentrosymmetric


MOFs  Nonlinear optics  Octupolar symmetry  Push–pull effect  Second harmonic
generation

Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
2 Metal–Organic Frameworks for Second-Order Nonlinear Optics . . . . . . . . . . . . . . . . . . . . . . . . . 148
2.1 NLO MOFs Built from Chiral Ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
2.2 NLO MOFs Built from Achiral Ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
3 Conclusions and Perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163

S. Du (*) and H. Zhang


Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fujian,
Fuzhou 350002, P. R. China
e-mail: swdu@fjirsm.ac.cn
146 S. Du and H. Zhang

Abbreviations

1,3-bimb 1,3-Bis(imidazol-1-ylmethyl)-benzene
1,4-bimb 1,4-Bis(imidazol-1-ylmethyl)-benzene
4-pya 4-Pyridylacrylate
aptz 5-(6-Aminopyridin-3-yl)tetrazol-1-ide
BCDC N,N0 -bis(4-cyanophenyl)-(1R,2R)-diaminocyclohexane
cda Carbamyldicyanomethanide anion
D-H2ca D-camphoric acid
DMF N,N-dimethylformamide
DPASD 4-(4-(Diphenylamino)styryl)-1-dodecylpyridinium
dpys 4,40 -Dipyridylsulfide
DSAT 4-N, N-dimethylamino-40 -N0 -methyl-stilbazolium tosylate
(E)-4-pyv-3-bza (E)-3-(2-(4-Pyridyl)vinyl)benzoate
(E)-4-pyv-4-bza (E)-4-(2-(4-pyridyl)vinyl)benzoate
H2dnty 3,5-Dinitrotyrosine
H2nicO 6-Hydroxynicotinic acid
H2oba 4,40 -Oxybis(benzoic acid)
H2SCMC S-carboxymethyl-L-cysteine
H2sdba 4,40 -Sulfonyldibenzoic acid
H3tzpbin (1R,2R)-1-(3-(1H-tetrazol-5-yl)phenyl)-2-(pyridin-4-yl)-
2,3-dihydro-1H-benzo[e]indole
HAmidn 2-Amino-4,5-imidazoledicarbonitrile
H-Imazethapyr (2-(4,5-Dihydro-4-methyl-4-(1-methylethyl)-5-oxo-1H-
imidazol-2-yl)-5-ethyl-3-pyridinecarboxylic acid
Htrtr 3-(1,2,4-Triazole-4-yl)-1H-1,2,4-triazole
Htzpbin 1-(3-(1H-tetrazol-5-yl)phenyl)-2-(pyridin-4-yl)-3H-benzo[e]
indole
inic Isonicotinate
m-H2bdc 1,3-Benzenedicarboxylic acid
nic Nicotinate
OH-m-H2bdc 5-Hydroxyisophthalic acid
phen 1,10-Phenanthroline
pyb 4-(4-Pyridyl)benzoate
quin-6-c Quinoline-6-carboxylate
spcp 4-Sulfanylmethyl-40 -phenylcarboxylate pyridine
THF Tetrahydrofuran
Metal–Organic Frameworks for Second-Order Nonlinear Optics 147

1 Introduction

The development of NLO materials can be traced back to the first observation of
second-harmonic generation (SHG) by Franken et al. in 1961 in a quartz crystal
when a ruby laser was used as incident light resource [1]. Since then, intensive
research has been carried out in the field of NLO materials as they can expand the
range of laser wavelengths. At present, NLO have been identified as excellent
candidates for emerging photonic technologies based on the fact that photons are
capable of processing information with the speed of light. They also possess vast
potential for use in a variety of photonic systems, including high-speed optical
modulators, ultra-fast optical switches, and high-density optical data storage media.
These devices are essential for continued advancement in the effort to transform
information storage and transmission from electrical to the optical regime [2–4].
In the beginning, most interest in NLO mateirals was focused on inorganic
materials, such as quartz, potassium dihydrogen phosphate (KDP), lithium niobate
(LiNbO3), potassium titanyl phosphate (KTP), β-barium borate and semiconductors
such as cadmium sulfide, selenium, and tellurium. Although many of them have been
widely used in commercial frequency conversion for lasers and optical parametric
generator, the choice of these materials is still rather limited [5, 6]. Due to the lack of
extended π-electron delocalization, purely inorganic materials usually have low NLO
responses (e.g., semiconductors). They also have some drawbacks such as difficulty
of synthesis, lack of optical quality, and slow electro-optic response times. Organic
compounds are other potential candidates for NLO materials because of their strong
NLO properties, high optical damage threshold, low cost and short response time to
optical excitation. Urea, for example, is one of the most promising materials for NLO
application, with SHG being 2.5 times that of ADP [7]. It has become evident that
conjugated molecules having a strongly electron-donating group (D) connected to an
electron-accepting unit (A) via a π-conjugated bridge often show highly nonlinear
optical effects due to the D!A intramolecular charge transfer (ICT) [8]. In fact,
some of them exhibit the largest known nonlinear coefficients, often considerably
larger than those of their inorganic counterparts. For example, the second-order NLO
coefficient of DSAT crystals is ten times as large as that of the inorganic standard
LiNbO3 [9]. The major advantage of using organic materials for NLO applications
lies in the ease of fine-tuning of the NLO properties by rational modification of
their molecular structure and functional groups. Moreover, organic compounds, in
general, are easy to process and integrate into optical devices. Unfortunately, most
of the organic π-conjugated molecules crystallize in centrosymmetric space groups
due to favored anti-parallel stacking, therefore producing materials with no second
order bulk susceptibility. They also suffer from poor physicochemical stability, low
hardness, and cleavage tendency that obstruct their device applications.
Metal–organic frameworks, often called MOFs, are a new class of inorganic–
organic hybrid materials that have enormous potential for many practical
applications [10]. They can combine the high nonlinear optical coefficients of the
organic molecules with excellent physical properties of the inorganics, and may
148 S. Du and H. Zhang

thus be promising candidates for NLO materials. The coordination of organic


ligands to metal ions can not only give rise to the metal-to-ligand charge transfer
transitions, which are important for SHG, but also allow the organic chromophores
to be arranged in orderly geometries, such as octahedral, tetrahedral and those that
are less commonly observed in organic materials, enabling multidirectional charge
transfer. Furthermore, a large number of such compounds possessing a range of
network structures can be readily made by choosing appropriate molecular building
blocks (typically metal ions or clusters) in combination with multitopic organic
ligands. The almost unlimited combination of building blocks and organic linkers
would, in principle, lead to assemble of a vast number of MOFs, allowing the
large-scale screening and design of new NLO-active MOF materials. In addition,
MOFs offer a large number of design possibilities and the advantage of tailorability:
a fine-tuning of the NLO properties can be achieved by rational modification of
the chemical composition and structure [11]. While a variety of materials including
inorganic, organometallic, organic, and polymeric systems have been studied for
their NLO activity, the development of NLO materials based on metal–organic
frameworks still remains in its infancy [12].

2 Metal–Organic Frameworks for Second-Order


Nonlinear Optics

For a material to exhibit NLO activity it should be noncentrosymmetric. Therefore,


rational design and synthesis of noncentrosymmetric MOFs is the prerequisite
and key issue to obtain NLO MOFs materials. A straightforward strategy for the
efficient synthesis of noncentrosymmetric MOFs includes the coordination of chiral
organic ligands to metal ions or the use of chiral templates to induce homochirality
into the framework [13]. They can also be made from achiral components via
spontaneous resolution during crystal growth without involving any enantiopure
precursors [14, 15]. In some cases, unsymmetrical ligands can help remove the
centrosymmetry from the frameworks, in particular those with diamondoid or grid
topologies [12]. Another synthetic approach to noncentrosymmetric MOFs is based
on mixed metal carboxylates. For example, it has been found that the incorporation
of alkali or alkaline earth ions into the Cd2+/carboxylate frameworks may greatly
enhance the possibility of generating acentric MOFs, despite that the underlying
causes of noncentrosymmetry have not been fully elucidated [16, 17]. By following
the approaches outlined above, a large number of noncentrosymmetric MOFs have
been successfully synthesized. However, in order to minimize optical losses in SHG
processes, only those MOFs which are transparent in the working frequency range
may have possible NLO applications. Thus, we will focus on the synthesis,
structures of MOFs with d10 metal ions such as Zn2+, Cd2+, and pay more attention
to those with large SHG efficiencies.
Metal–Organic Frameworks for Second-Order Nonlinear Optics 149

2.1 NLO MOFs Built from Chiral Ligands

The chirality associated with MOFs can ensure noncentrosymmetric solid-state


organization, which suggests potential second-order nonlinear optical effects in
these materials. In principle, all chiral MOFs are SHG active. However, they do
not necessarily display strong SHG responses as the dipoles of chromophoric
building blocks in the chiral MOFs structures do not have to orient in a noncentro-
symmetric fashion. Although it has been found that the mode of metal coordination
and MOF topology may influence the solid-state SHG in these chiral MOFs,
the direct correlation between the structures and the observed SHG has not yet
been established. Nevertheless, it has become evident that, similar to organic NLO
materials, chiral MOFs with push–pull centers (with electron donor–acceptor
systems) generally have large SHG responses [18].
The design and construction of chiral MOFs can be readily achieved by self-
assembly of metal ions with chiral ligands, which involves the transfer of chiral
information from a chiral ligand to the metal center and then to the whole frame-
work. It should be pointed out that this is not the surest way to produce chiral
MOFs, as in some cases mesomeric phenomenon may occur and the path for the
transfer of chiral information may be interrupted. In this context, chiral ligands can
be classified into two types: nonaromatic and aromatic chiral ligands in terms of
whether they contain a conjugated system or not. Nonaromatic chiral ligands, in
particular those with carboxylic groups and/or N-donor atoms have the capability to
organize metal ions into chiral or acentric MOFs. For example, the reactions of
S-carboxymethyl-L-cysteine with Zn2+ and Cd2+ ions gave two chiral MOFs
[Zn(H2O)(SCMC)]n and {[Cd(H2O)(SCMC)] · 2H2O}n [19], which exhibit wave-
like 2D (4,4) layers that are further assembled into a 3D framework through
hydrogen bonding interactions. Both of them crystallize in the chiral space groups
and show SHG intensities of 0.05 and 0.06, respectively relative to that of urea.
A 3D chiral MOF [Mn(H2O)(SCMC)] constructed by Mn2+ and SCMC was also
synthesized [20]. Structural analysis reveals that the four symmetry-related Mn2+
ions in this compound are bridged by four SCMC ligands through N and O atoms,
generating a 3D MOF with a 66 topology. Upon excitation by a high-intensity
Nd:YAG laser, a weak green light emission from the crystal was observed,
confirming its noncentrosymmetric structure.
Some other naturally chiral carboxylate ligands were also selected to construct
chiral MOFs. For example, assemblies of Zn2+ and Cd2+ with D-camphoric acid in
combination with some semi-rigid N-donor auxiliary ligands, such as 1,3- and
1,4-bis(imidazol-1-ylmethyl)benzene and 4,40 -dipyridylsulfide, result in the forma-
tion of chiral MOFs [M(1,3-bimb)(D-ca)]n (M ¼ Zn, Cd), [Zn(1,4-bimb)(D-ca)]n
and a noncentrosymmetric MOF {[Cd2(dpys)(D-ca)2(H2O)2] · H2O}n, in which the
1D Zn2+(Cd2+)/D-ca chains are linked by N-donor ligands to form either a 2D
network or a ladder-like structure [21]. Preliminary experimental results indicate
that they exhibit SHG intensities ca. 0.3, 0.1, 0.3, and 0.8 times, respectively,
as large as that of urea. In fact, many MOFs with nonaromatic chiral ligands are
150 S. Du and H. Zhang

Scheme 1 (a) The donor-π-acceptor system in a ligand. (b) Schematic drawing of L showing the
two-center-D–A system. (c) Schematic drawing of H3tzpbin

Fig. 1 (a) The 2D network of [Zn(tzpbin)2] · 1.5H2O. After [23]. (b) The 1D infinite molecular box
of [Cd(H2tzpbin)2(H2O)2]n. After [24]

NLO active. However, they usually have weak SHG efficiencies, probably due to
the lack of push–pull effect centers, an essential criterion for an SHG response.
Basically, push–pull electronic effect is more easily achieved with an aromatic
chiral ligand than with a nonaromatic one as the former contains a π-conjugated
system (Scheme 1a). Some chiral ligands with a push–pull character have been thus
designed and employed in the reaction of metal ions, forming a range of MOFs that
exhibit good SHG effect. For instance, a chiral ligand, 3-((1R,2R)-2-(pyridin-4-yl)-
2,3-dihydro-1H-benzo[e]indol-1-yl)benzonitrile (L) with two chiral centers and a
two-center-D–A (donor–acceptor) system has been designed (Scheme 1b) [22].
Under hydrothermal conditions, its cyano group can be readily converted into a
tetrazole functionality in the presence of NaN3, in situ generated a chiral tetrazole
ligand H3tzpbin (Scheme 1c), which subsequently reacted with Zn2+ and Cd2+ ions
to afford a 2D square grid network [Zn(tzpbin)2] · 1.5H2O (Fig. 1a) and a 1D infinite
molecular box [Cd(H2tzpbin)2(H2O)2]n (Fig. 1b) [23, 24]. It should be pointed out
that the reaction to [Zn(tzpbin)2] · 1.5H2O was accompanied by a dehydrogenation
process, making the ligand nonchiral. As the result, the product crystallized in a
nonchiral space group (Fdd2). Due to the existence of a two-center-D–A system
Metal–Organic Frameworks for Second-Order Nonlinear Optics 151

Scheme 2 Schematic drawing of H2dnty and BCDC ligands

(multicenter push–pull electronic effect), L displays a strong SHG response, approx-


imately 12 times larger than that of urea. In contrast, H3tzpbin has weak SHG
effect, nearly the same as that of urea. This could be due to the fact that different
from 4-cyanophenyl group in L, the tetrazole ring in H3tzpbin is not an electronic
acceptor. However, when it coordinates with metal ions, donation of the lone pair
of electrons from pyridyl and tetrazole nitrogen atoms onto metal center occurs,
creating an excellent donor–acceptor system in [Zn(tzpbin)2] · 1.5H2O and
[Cd(H2tzpbin)2(H2O)2]n. As a result, they both exhibit very strong SHG responses,
about 50 and 80 times, respectively, that observed for urea.
As one can see above, MOFs with a donor-π-acceptor organic chromophore
ligand may have good SHG responses as the hybridization of organic chromophore
and metal ions favors charge transfer in general. This strategy is not limited to closed-
shell d10 metals. An Mn2+ MOF [Mn(Hdnty)2] constructed by 3,5-dinitrotyrosine
(Scheme 2a) has been shown to have a powder SHG efficiency of six times that of
urea [25]. This compound crystallizes in a chiral space group P212121 and has a 2D
square-grid topological net. Although the homochiral H2dnty chromophore itself
displays a weak SHG response, once it is combined with Mn2+ ions, the resulting
hybridized MOF can produce a large SHG effect. It appears that the coordination
with metal ions may enhance the push–pull effect of the ligand, even though some
adverse impact has been observed. A lanthanide compound, [Nd(Hdnty)2(NO3)
(H2O)5] · 3H2O synthesized using the same ligand has also been reported [25].
It is a 3D framework assembled through weak supramolecular interactions, such as
π–π stacking and hydrogen bonds between amino and nitro groups. Similarly, this
compound exhibits a strong SHG effect with an intensity of five times as great
as that of urea. In this case, however, the enhancement of SHG efficiency is
partially attributed to the hydrogen bonding that can also be considered as a good
donor–acceptor system capable to increase charge separation.
Another interesting chiral ligand BCDC with multiple ligation sites and a
push–pull center (Scheme 2b) was applied in the synthesis of NLO MOFs. Reaction
of BCDC with Cu+ and Ag+ salts yielded chiral MOFs [Cu(BCDC)]PF6 · THF and
[Ag(BCDC)]ClO4 [26, 27]. The copper compound is a 2D network in which the
Cu+ ions reside on a twofold axis in the crystal structure and adopt a tetrahedral
environment by coordinating with two amino groups of a BCDC ligand and cyano
groups of two other BCDC ligands. The silver compound has a 1D helical chain
structure in which the Ag+ ion is coordinated by two cyano nitrogen atoms from
152 S. Du and H. Zhang

two different BCDC ligands in a linear coordination geometry. The SHG values
observed for BCDC, [Cu(BCDC)]PF6 · THF, and [Ag(BCDC)]ClO4 are about
0.3, 0.2 and 2.9 times, respectively, as large as that of urea. This trend is consistent
with traditional push–pull concepts and closely related to different configurations
and local coordination environment of BCDC ligand in the crystal structure,
which is further supported by theoretical calculations. It follows that in addition
to push–pull effect, the molecular nonlinearity may also exert a great influence
on the bulk NLO property.

2.2 NLO MOFs Built from Achiral Ligands

Despite the tremendous success in the construction of NLO MOFs from chiral
ligands, this synthetic strategy has suffered from the high cost and limited avail-
ability of the enantiopure chiral ligands. Thus, an alternative method by using
achiral ligands is required for the development of such materials. Some chiral
MOFs can be obtained by spontaneous resolution without involving any chiral
components. Unfortunately, the formation of chiral products through spontaneous
resolution is still unpredictable because the laws of physics determining this
processes remain poorly understood up until now. Another simple but effective
way to increase the chance of making noncentrosymmetric MOFs is to use unsym-
metrical bridging ligands. The unsymmetrical ligands, in particular those with
strong donor/acceptor substituents may, on the one hand, help remove the center
of symmetry and, on the other hand, create electronic asymmetry in the MOFs
network, which is the key for a high nonlinearity or substantial SHG response.
Following this principle, a range of noncentrosymmetric MOFs, with 1D, 2D,
and 3D network structures have been synthesized. Their SHG efficiencies vary in
a wide range and appear to have no direct correlation with the types of metal ions,
the ligands and the topological structures. However, it has been observed that
MOFs with a 3D diamondoid structure similar to that of KDP, or with an octupolar
symmetry usually exhibit large SHG activity. In addition, some 2D grids and 1D
helical chains also display good SHG efficiencies.

2.2.1 NLO MOFs with 3D Diamondoid Structures

The approach to the rational design and synthesis of noncentrosymmetric MOFs by


using unsymmetrical bridging ligands work best for the 3D diamondoid system.
The pioneering work in this field was carried out by Lin’s group. They demonstrate
that a centrosymmetry breaking at a diamond net can happen when the metal
tetrahedral nodes of the diamond net are bridged by unsymmetrical ligands. They
first synthesized a chiral threefold interpenetrated diamondoid net, [Zn(inic)2]
from Zn(ClO4)2 · 6H2O and inic anion (generated in situ from the hydrolysis of
4-cyanopyridine) under solvothermal conditions (Fig. 2). The tetrahedral Zn2+
Metal–Organic Frameworks for Second-Order Nonlinear Optics 153

Fig. 2 A 3D diamondoid net of [Zn(inic)2] built from Zn2+ tetrahedral connecting points and
p-pyridinecarboxylate linkers. After [12]

Scheme 3 3D MOFs of diamondoid topology built from Zn2+/Cd2+ tetrahedral connecting nodes
and p-pyridinecarboxylate type of unsymmetrical ligands

centers in this compound are connected through pyridine and carboxylate groups of
unsymmetrical inic anions to form a noncentrosymmetric threefold interpenetrated
diamondoid framework [28]. This compound shows a weak SHG signal, being 1.5
times more intense than that of α-quartz (Scheme 3a).
By tuning the length of p-pyridinecarboxylate linkers, similar reactions between
Zn(ClO4)2 · 6H2O or Cd(ClO4)2 · 6H2O and corresponding precursors afforded more
noncentrosymmetric diamondoid MOFs with systematically elongated linear spacers.
It has been found that the longer the bridging ligand, the higher the interpene-
tration degree. For example, the reactions with 4-pyridylacrylic acid led to
diamondoid networks, [Zn(4-pya)2] [28], and [Cd(4-pya)2] · H2O [29] with fivefold
interpenetration (Scheme 3b), while with the longer 4-(4-pyridyl)benzoate and
(E)-4-(2-(4-pyridyl)vinyl)benzoate afforded a sevenfold interpenetrated [Cd(pyb)2] ·
H2O (Scheme 3c) and eightfold interpenetrated [M((E)-4-pyv-4-bza)2] (M ¼ Zn, Cd)
(Scheme 3d) diamondoid networks, respectively [29, 30]. As expected, all these
154 S. Du and H. Zhang

Fig. 3 (a) Asymmetric unit representation of [Zn(aptz)2] showing that the Zn center displays
a slightly distorted tetrahedron. After [23]. (b) A non-interpenetration diamondoid net of
[Zn(aptz)2]. After [23]

compounds crystallize in the noncentrosymmetric space group. This series of


diamondoid MOFs exhibit impressive SHG activities, ranging from 1.5 to
400 times that of α-quartz. In particular, the SHG properties of [M((E)-4-pyv-4-bza)2]
are comparable to that of the technologically important lithium niobate material.
Furthermore, as the length of bridging ligand gets longer, the SHG efficiency of
the corresponding diamondoid MOF increases, a consequence of better conjugation
between the donor and the acceptor, which enhances the molecular hyperpolari-
zability β of NLO chromophores. This work demonstrates for the first time the
possibility of fine-tuning the NLO properties of diamondoid MOFs through simple
change of the lengths of ligand spacers.
In addition to choosing a long unsymmetrical bridging ligand, the combination
of a diamondoid structure and a good push–pull system of ligand can also effec-
tively improve the SHG properties. For example, Xiong and co-workers [23]
synthesized a non-interpenetrated diamondoid Zn-tetrazole MOF [Zn(aptz)2]
which showed an SHG activity ca. 5 times that of urea. In this compound, the
Zn2+ center adopts slightly distorted tetrahedral coordination geometry, surrounded
by four N atoms from four aptz ligands (Fig. 3a). The larger steric bulk of the aptz
ligand together with the short separation between Zn2+ ions prevents the interpene-
tration of the individual diamondoid networks (Fig. 3b). The strong SHG efficiency
of this compound is attributed to the presence of strong donor/acceptor substituents
in the aptz ligand.
Besides unsymmetrical ligands, racemic chiral ligands are also useful for
making noncentrosymmetric diamondoid MOFs. Racemic H-Imazethapyr ligand,
for example, has been used to react with Cd(ClO4)2 · 6H2O under hydrothermal
conditions to give a diamondoid MOF, [Cd(Imazethapyr)2] [31]. The Cd2+ ions in
this MOF adopt a slightly distorted octahedral geometry, coordinated with four N
atoms and two O atoms from Imazethapyr anions (Fig. 4a). Thus, each Cd2+ ion
links four different Imazethapyr ligands in four different directions, generating a 3D
framework with a diamondoid topology. No interpenetration occurs between the
individual diamondoid networks due to the steric effect of the ligand and the short
Metal–Organic Frameworks for Second-Order Nonlinear Optics 155

Fig. 4 (a) Coordination environment of Cd2+ in [Cd(Imazethapyr)2]. After [31]. (b) A simplified
3D diamondoid net of [Cd(Imazethapyr)2]. After [31]

distance between two adjacent Cd2+ ions (Fig. 4b). Preliminary studies of a powdered
sample indicate that this compound is SHG active with a value ca. 20 times
larger than that of KDP.

2.2.2 NLO MOFs with Octupolar Symmetries

The concept of octupolar nonlinearities was described in the early 1990 by Zyss
and co-workers [32]. Due to the lack of ground state dipole moment, which is the
main origin of centrosymmetric molecular packing style, octupolar chromophores
show easier noncentrosymmetric crystallization in the solid state to achieve a
maximum bulk effect. Besides, octupolar frameworks can be more transparent
and the efficiency-transparency trade-off can be remarkably improved. Because
of these advantages, increasing research attention has been focused on NLO active
octupolar molecules, including organic compounds and metal coordination and
organometallic complexes. Although octupolar MOFs are rarely encountered in
MOFs, some of them have indeed proven to have excellent NLO properties.
A synthetic approach to noncentrosymmetric octupolar NLO materials involves
using trinuclear metal carboxylates clusters as fundamental building units. In general,
trinuclear metal carboxylates of group 12 metals are more difficult to obtain in
comparison with those of other transition metals. Despite this, an octupolar cadmium-
carboxylate MOF, formulated as [Cd3(μ-OH)3((E)-4-pyv-4-bza)6(py)6](ClO4)2, was
successfully synthesized from Cd(ClO4)2 · 6H2O and ethyl 4-[2-(4-pyridyl)ethenyl]
benzoate under hydro(solvo)thermal conditions [33]. The trinuclear hydroxo-
centered Cd2+ carboxylates, acting as a secondary building unit (SBU), possesses
a threefold rotational symmetry and constitutes the octupolar NLO chromophoric
unit of this compound (Fig. 5a). The SBUs are then linked via ligands to form a
chiral 2D layer structure (Fig. 5b). Kurtz powder SHG measurements for this
compound revealed a powder SHG intensity of 130 versus α-quartz.
156 S. Du and H. Zhang

Fig. 5 (a) View of the cationic octupolar building block in [Cd3(μ-OH)3((E)-4-pyv-4-bza)6(py)6]


(ClO4)2. After [33]. (b) Schematic showing of the 2D sheet down from the c axis. After [33]

Fig. 6 (a) View of the hexameric octupolar building block [Cd6(C2O4)8]4. After [34]. (b) View
of the octupolar 3D anionic open framework of (H2NMe2)2[Cd3(C2O4)4] · MeOH · 2H2O.
After [34]

Another octupolar cadmium MOF with a large NLO efficiency (150  α-quartz)
has also been synthesized by a solvothermal reaction of CdCO3 with symmetrical
oxalic acid [34]. The product (H2NMe2)2[Cd3(C2O4)4] · MeOH · 2H2O isolated
from this reaction crystallizes in the cubic acentric space group and possesses a 3D
porous coordination network. Two sets of equilateral trianglular [Cd3(C2O4)4]2
motifs with the center being located at a threefold axis form a bowl-shaped
hexameric basic building block [Cd6(C2O4)8]4 (Fig. 6a). Face-to-face connection
of the Cd3 triangles forms Cd9 cages, which undergo face-to-face intercage packing,
leading to an octupolar 3D anionic supercage network (Fig. 6b) with the small
and large cavities being occupied by solvent molecules and H2NMe2+ cations,
respectively. Complete replacement of H2NMe2+ by NH4+, Na+, and K+ was
accomplished and the resulting MOFs have powder SHG intensities of 155, 90,
and 110, respectively, versus α-quartz, showing interesting cation-dependent SHG
responses.
Metal–Organic Frameworks for Second-Order Nonlinear Optics 157

Fig. 7 (a) View of 2D square grid structure of [Zn(nic)2]. After [35]. (b) A schematic showing of
the interweaving of three independent rhombohedral grids in [Cd((E)-4-pyv-3-bza)2]. After [36].
(c) ABC stacking of four independent 2D rhombohedral grids in [Zn4((E)-4-pyv-3-bza)8] ·
(H(E)-4-pyv-3-bza) · H2O. After [36]

2.2.3 NLO MOFs with 2D Grid Networks

As discussed above, assemblies of Zn2+/Cd2+ ions with unsymmetrical linear


bridging ligands under hydro(solvo)thermal conditions permit the fabrication of
diamondoid structures. Unfortunately, this method is not guaranteed to produce
diamondoid MOFs as MOF topology is highly sensitive to small changes in the
reaction conditions or the organic ligands. Sometimes 2D grid-type MOFs can
be formed. In these cases, however, the use of unsymmetrical ligands can still
increase the chance of obtaining noncentrosymmetric MOFs. For example, by using
m-pyridinecarboxylate linkers instead of p-pyridinecarboxylate ligands, a couple
of 2D grids were obtained. Similar to the diamondoid system, the elongation
of ligand length results in better conjugation between the donor and the acceptor,
and hence increases the NLO activity. Thus, the hydro(solvo)thermal reaction
of zinc perchlorate hexahydrate with 3-cyanopyridine gave a 2D square grid
[Zn(nic)2], in which the Zn center has a highly distorted octahedral geometry
with two carboxylate groups and two pyridyl nitrogen atoms of four different
nicotinate groups in a cis configuration (Fig. 7a). This compound shows a low
SHG effect (2  α-quartz) [35]. Similar reactions of Cd2+ and Zn2+ salts with a
longer bridging ligand 4-(3-cyanostyryl)pyridine yielded 2D rhombohedral grids
[Cd((E)-4-pyv-3-bza)2] and [Zn4((E)-4-pyv-3-bza)8] · (H(E)-4-pyv-3-bza) · H2O,
respectively [35, 36]. The former has a interweaving of three independent
rhombohedral grids in the ac plane (Fig. 7b) and exhibits higher SHG intensity
(1,000  α-quartz) than the technologically important LiNbO3, whereas the latter
has an ABC stacking fashion of four independent 2D rhombohedral grids along
the c axis (Fig. 7c) and also display a high SHG intensity (400  α-quartz).
158 S. Du and H. Zhang

Fig. 8 (a) The 2D layer structure of [Cd8(nicO)8(phen)8] · H2O (phen is omitted for clarity). After
[37]. (b) The 2D layer structure of [Eu(cda)3(H2O)3] · H2O. After [38]

By using mixed ligands of unsymmetrical 6-hydroxynicotinic acid and


1,10-phenanthroline, a noncentrosymmetric grid-like MOF [Cd8(nicO)8(phen)8] ·
H2O that showed impressive SHG activity was synthesized [37]. Two distinct
Cd centers, one is in a distorted mono-capped triangular prismatic coordination
geometry and the other in a distorted octahedral geometry, form dinuclear building
blocks which are further linked via nicO ligands to form rhombus grids extending
along the ab plane (Fig. 8a). SHG measurement on the powder sample reveals
that it exhibits an SHG efficiency of ca. 8 times that of KDP. Lanthanide
metals are also capable to form 2D grids with polynitrile ligands. For example, a
noncentrosymmetric 2D polynitrile Eu3+ MOF [Eu(cda)3(H2O)3] · H2O has been
synthesized. The Eu3+ center is eight-coordinate with three nitrogen atoms from
nitrile groups and five oxygen atoms: three are from water molecules and two are
from amide groups (Fig. 8b). The coordination of amide-oxygen atoms to Eu3+
suggests a strong conjugation in cda. As a result, this compound shows strong
powder SHG efficiency, being 16.8 times that of urea [38].

2.2.4 NLO MOFs with 1D Helical Chains

The systematical design of noncentrosymmetric MOFs based on 1D chains is more


difficult compared to those with high-dimensional structural motifs because the 1D
chain systems lack control in two other directions. Except when using enantiopure
chiral ligands, most of the noncentrosymmetric linear chains (1D MOFs) are
synthesized accidentally and normally have a weak SHG activity. However, some
2D and 3D MOFs constructed by helical chain motifs can display good SHG responses.
A 2D zinc MOF [Zn(spcp)(OH)] based on two types of homochiral helices
was synthesized through spontaneous resolution [39]. The ligand is a sulfur-
containing asymmetrical linker where sulfide moieties are well known as redox
active functional groups that could enhance electronic asymmetry. The structural
feature of this MOF is the alternating arrangement of the two distinct homochiral
helices along the crystallographic b axis (Fig. 9a). One helix is formed by hydroxo
bridging Zn2+ atoms while the other with an opposite handedness is constructed
Metal–Organic Frameworks for Second-Order Nonlinear Optics 159

Fig. 9 (a) View of the two types of homochiral helices in [Zn(spcp)(OH)]. After [39].
(b) View of the 3D net of [Zn(OH)(quin-6-c)] along the c axis. After [40]

by spcp bridged between the Zn2+ centers. The compound crystallizes in a P21
space group and the preliminary experimental results show that it displays high
powder SHG efficiency approximately five times higher than that of KDP. The
large SHG response could be attributed to the big electronic asymmetry of infinite
hydroxo-zinc helices. Another zinc MOF [Zn(OH)(quin-6-c)] containing similar
hydroxo-zinc helices has been reported [40]. Each [Zn(μ-OH)2]n helix in this
compound is linked to four nearest-neighbors through the bridging quin-6-c
ligands along the a and b axes, generating a 3D 4-connected framework (Fig. 9b).
This compound crystallizes in a noncentrosymmetric orthorhombic space group
and displays a powder SHG intensity ca. 460 times in comparing with that of
α-quartz.
1,2,4-Triazole and its derivatives have been widely used to design MOFs.
These π-electron-rich heterocyclic ligands with versatile coordination modes are
found to have great promise for constructing NLO MOFs. A typical example [41]
of these compounds is [Cd(trtr)2]n with a noncentrosymmetric polar packing
arrangement. The basic structural motif of this compound is a helical chain made
up on trtr ligands and Cd atoms. These right- and left-handed helices are alterna-
tively arranged, leading to a 3D architecture that can be reduced to a binodal
structure with six-connected Cd nodes and three-connected trtr nodes. It crystallizes
in a noncentrosymmetric space group (Fdd2) and displays an SHG signal six times
that of KDP. The strong SHG response is mainly attributed to the good donor–
acceptor system of the organic moieties which are arranged in an orderly fashion
(Fig. 10a). Besides, due to the packing of the helical chains and sheets in parallel,
the whole 3D crystal structure also has a dipole moment along the c direction which
is believed to strengthen the hyperpolarizability, giving rise to a good SHG
response.
Imidazole derivatives have been used with metal ions to form noncentro-
symmetric MOFs as they have strong tendency to induce helical structures of axial
chirality. The im-NM+ donor–acceptor systems in these MOFs are expected
to enhance the hyperpolarizability β and thus the second-order NLO susceptibility.
By introducing polar substituents such as – CN into the imidazole ring, the β value
160 S. Du and H. Zhang

Fig. 10 (a) View of the 3D structure of [Cd(trtr)2]n made up parallel right-helical chains
(the red arrows represent dipoles of trtr ligands). After [41]. (b) View of the triple helices in
[Ag(Amidn)]n. After [42]. (c) Schematic drawings of Htrtr and HAmidn ligands

can be further increased. This is well demonstrated by the rigid planar imidazole
ligand HAmidn which is able to combine with Ag+ ions to form a noncentro-
symmetric 3D MOF [Ag(Amidn)]n having good SHG efficiency[42]. The structure
features nanotubes built of chiral Ag-Amidn triple helices which are believed
to be the source of the resulting noncentrosymmetric space group (Fig. 10b). The
triple right- and left-handed helices are alternately arranged to form a helix-based
open framework structure. The SHG efficiency of this compound is approximately
5.4 times that of KDP. The strong SHG activity is contributed to the donor–acceptor
system in which the Ag+ ion acts as an electron donor while the imidazole ring
as an electron acceptor.

2.2.5 NLO MOFs Incorporating with Alkali or Alkaline Earth Ions

In MOF synthesis, alkali or alkaline earth ions are inclined to form a cation
hydration complex, acting only as a counter ion when the reaction is carried
out in aqueous solution, as usually happens in the hydrothermal synthesis where
water is used as a reaction medium. However, in the absence of water, they have a
great opportunity to coordinate with carboxylate oxygen atoms and participate
in the network construction. A number of Cd2+/Li+ carboxylate MOFs have thus
been synthesized via solvothermal reactions of Cd2+ salt with various symmetrical
aromatic polycarboxylic acids in the presence of LiNO3. The main feature of
these compounds is that they are all anionic frameworks in which the Cd2+ center
is eight-coordinate, chelated by four carboxylate groups in a severely distorted
mono-capped pentagonal bipyramidal geometry while the Li+ ion acts as both a
charge balancer and a connecting node to extend the network structure. The eight-
coordinated Cd2+ centers are different from those commonly observed in the
centrosymmetric Cd2+/carboxylate frameworks where the octahedral Cd2+ centers
are chelated by two carboxylate groups, leaving two mutually cis coordination sites
to be occupied by solvent molecules or nitrogen atoms from the auxiliary ligands.
Metal–Organic Frameworks for Second-Order Nonlinear Optics 161

Fig. 11 (a) Polyhedral view of the 3D anionic framework of [CdLi(oba)2]. (b) View of the
Li+-omitted fivefold interpenetrating diamondoid framework. (c) The catenation of the
six-membered rings of five individual diamondoid nets. Reprinted with permission from [16]
Copyright 2010 American Chemical Society

Despite the symmetrical ligands used, the above mixed metal system has a strong
tendency to afford noncentrosymmetric structures [16], presumably due to the pres-
ence of mixed metal connecting nodes that favors noncentrosymmetric packing.
For example, two noncentrosymmetric Cd2+/Li+ MOFs, (Me2NH2)[CdLi(oba)2]
and (Me2NH2)[CdLi(m-bdc)2] were synthesized this way. In (Me2NH2)[CdLi(oba)2],
the Cd2+ center is in a distorted mono-capped pentagonal bipyramidal geometry,
chelated by four carboxylate groups while the Li+ ion is in a slightly deviated
tetrahedral geometry, coordinated with four O atoms from four different carboxy-
late groups. The Cd2+ and Li+ ions are connected through oba2 ligands to form a
3D anionic framework (Fig. 11a). If the Li+ ions are ignored, such a 3D framework
can be abstracted as a typical diamondoid topological network with the Cd–Cd–Cd
angles of 109.42 and 109.57 , being close to that of the perfect tetrahedral
angle found in diamond (Fig. 11b). The resulting Li+-omitted framework is a
fivefold self-interpenetrated net where the six-membered rings of one diamondoid
net is catenated by four six-membered rings of the other identical diamondoid nets
(Fig. 11c). Thus, the overall framework of this compound can be considered as
fivefold interpenetrating diamondoid nets which are linked by Li+ connections.
The coordination environment of Cd2+ and Li+ ions in (Me2NH2)[CdLi(m-bdc)2]
is similar to that in the oba2 compound (Fig. 12a). The interconnections of
m-bdc2 ligands and Cd2+, together with Li+ ions, result in a 3D anionic framework
with large square channels accommodated by the (Me2NH2)+ cations (Fig. 12b).
Upon removing the Li+ ions from the structure, the remaining anionic framework
becomes a 2D square grid net with a (4,4) topology (Fig. 12c). Therefore, this
compound can be viewed as being constructed by the 2D grid nets through the
Cd–Li–Cd linkages. SHG measurements show that the oba2 and m-bdc2 MOFs
possess SHG intensities of five and four times that of the KDP standard, respec-
tively. The former compound gives a slightly larger SHG intensity than the
latter one, probably due to its diamondoid structure. Following this strategy, several
other noncentrosymmetric Cd2+/Li+ MOFs, such as (Me2NH2)[Cd(OH-m-Hbdc)
(OH-m-bdc)] · DMF · CH3OH · H2O and (Me2NH2)[CdLi(sdba)2] · DMF · CH3OH
have been prepared. By using alkaline earth ions instead of Li+, a series of
162 S. Du and H. Zhang

Fig. 12 (a) The coordination environment of Cd2+ and Li+ in (Me2NH2)[CdLi(m-bdc)2].


(b) Polyhedral view of the 3D anionic framework of [CdLi(m-bdc)2]. (c) View of the Li+-omitted
square grid in (Me2NH2)[CdLi(m-bdc)2]. Reprinted with permission from [16] Copyright 2010
American Chemical Society

Cd2+/M2+(M ¼ Ca, Sr) noncentrosymmetric MOFs, i.e. [CdCa(m-bdc)2(DMF)2],


[CdCa(OH-m-bdc)2(H2O)2] · 2Me2NH and [CdSr2(m-bdc)2(NO3)2(DMF)4] were
obtained [17]. However, they all display weak SHG efficiencies.

3 Conclusions and Perspectives

In this chapter we briefly review the recent progresses in the design and synthesis
of metal–organic frameworks as a new class of materials with potential for NLO
application. The noncentrosymmetric MOFs, discussed herein are synthesized by
using either chiral ligands or unsymmetrical ligands to assemble with suitable
metal ions under hydro(solvo)thermal conditions. Additionally, incorporation of
alkali or alkaline earth ions into Cd2+/carboxylate frameworks is also an effective
way to produce noncentrosymmetric MOFs.
Thanks to the strong and highly directional metal-ligand coordination bonds,
as well as the virtually unlimited structural possibilities involving combinations
of inorganic and organic building components, MOFs can be rationally designed
and structurally tailored to fit NLO applications by judiciously choosing metal
nodes and bridging organic ligands. It has been found that similar to NLO organic
materials, NLO MOFs with push–pull systems usually exhibit good SHG responses.
From the structural point of view, diamondoid or octupolar MOFs often show
impressive SHG efficiencies. However, a more reliable method to synthesize non-
centrosymmetric MOFs and a comprehensive theory for better understanding
and revealing the structure-NLO property relationship in the NLO MOFs are still
needed. More recently, inclusion of the ordered organic dipolar chromophore
DPASD in a microporous MOF has been reported [43]. This encapsulated MOF
shows a highly NLO activity, which opens up new possibilities in the area of
NLO MOF materials.
Metal–Organic Frameworks for Second-Order Nonlinear Optics 163

Acknowledgments We acknowledge the financial supports from the National Basic Research
Program of China (973 Program, 2012CB821702), the National Natural Science Foundation of
China (21233009 and 21173221) and the State Key Laboratory of Structural Chemistry, Fujian
Institute of Research on the Structure of Matter, Chinese Academy of Sciences.

References

1. Franken PA, Hill AE, Peters CW, Weinreich G (1961) Generation of optical harmonics.
Phys Rev Lett 7:118–119
2. Saleh BEA, Teich MC (1991) Fundamentals of photonics. Wiley, New York
3. Boyd RW (1992) Nonlinear optics. Academic, San Diego
4. Zyss J (1994) Molecular nonlinear optics, materials, physics and devices. Academic, New York
5. Fan YX, Eckardt RC, Byer RL, Route RK, Feigelson RS (1984) AgGaS2 infrared parametric
oscillator. Appl Phys Lett 45:313–315
6. Glass AM (1988) Materials for photonic switching and Information processing. MRS Bull
13:16–20
7. Donaldson WR, Tang CL (1984) Urea optical parametric oscillator. Appl Phys Lett 44:25–27
8. Davydov BL, Derkacheva LD, Dunina VV, Zhabotinskii ME, Zolin VF, Koreneva LG,
Samokhina MA (1970) Connection between charge transfer and laser second harmonic
generation. JETP Lett 12:16–19
9. Marder SR, Perry JW, Schaefer WP (1989) Synthesis of organic salts with large second-order
optical nonlinearities. Science 245:626–628
10. Farrusseng D (2011) Metal–organic frameworks: applications from catalysis to gas storage.
Wiley-VCH Verlag & Co. KgaA, Weinheim
11. Evans OR, Lin WB (2002) Crystal engineering of NLO materials based on metal–organic
coordination networks. Acc Chem Res 35:511–522
12. Wang C, Zhang T, Lin WB (2012) Rational synthesis of noncentrosymmetric metal–organic
frameworks for second-order nonlinear optics. Chem Rev 112:1084–1104
13. Nickerl G, Henschel A, Grünker R, Gedrich K, Kaskel S (2011) Chiral metal–organic
frameworks and their application in asymmetric catalysis and stereoselective separation.
Chemie Ingenieur Technik 83:90–103
14. Lin WB (2005) Homochiral porous metal–organic framework: why and how? J Solid State
Chem 178:2486–2409
15. Morris RE, Bu X (2010) Induction of chiral porous solids containing only achiral building
blocks. Nat Chem 2:353–361
16. Lin JD, Long XF, Lin P, Du SW (2010) A series of cation-templated, polycarboxylate-based
Cd(II) or Cd(II)/Li(I) frameworks with second-order nonlinear optical and ferroelectric
properties. Cryst Growth Des 10:146–157
17. Lin JD, Wu ST, Li ZH, Du SW (2010) Syntheses, topological analyses, and NLO-active
properties of new Cd(II)/M(II) (M ¼ Ca, Sr) metal–organic frameworks based on
R-isophthalic acids (R ¼ H, OH, and t-Bu). Dalton Trans 39:10719–10728
18. Zhao H, Qu ZR, Ye HY, Xiong RG (2008) In situ hydrothermal synthesis of tetrazole
coordination polymers with interesting physical properties. Chem Soc Rev 37:84–100
19. Wang YT, Fan HH, Wang HZ, Chen XM (2005) Homochiral helical wavelike (4,4) networks
constructed by divalent metal ions and S-carboxymethyl-L-cysteine. J Mol Struct 740:61–67
20. Xu W, Liu W, Yao FY, Zheng YQ (2011) Synthesis, crystal structure and properties of the
novel chiral 3D coordination polymer with S-carboxymethyl-L-cysteine. Inorg Chim Acta
365:297–301
21. Liang XQ, Li DP, Li CH, Zhou XH, Li YZ, Zuo JL, You XZ (2010) Syntheses, structures, and
physical properties of camphorate coordination polymers controlled by semirigid auxiliary
164 S. Du and H. Zhang

ligands with variable coordination positions and conformations. Cryst Growth Des 10:
2596–2605
22. Zhao H, Li YH, Wang XS, Qu ZR, Wang LZ, Xiong RG, Abrahams BF, Xue Z (2004)
Noncentrosymmetric organic solids with very strong harmonic generation response. Chem
Eur J 10:2386–2390
23. Ye Q, Li YH, Song YM, Huang XF, Xiong RG, Xue Z (2005) A second-order nonlinear optical
material prepared through in situ hydrothermal ligand synthesis. Inorg Chem 44:3618–3625
24. Ye Q, Tang YZ, Wang XS, Xiong RG (2005) Strong enhancement of second-harmonic
generation (SHG) response through multi-chiral centers and metal-coordination. Dalton
Trans 1570–1573
25. Ye Q, Li YH, Wu Q, Song YM, Wang JX, Zhao H, Xiong RG, Xue Z (2005) The first
metal (Nd3+, Mn2+, and Pb2+) coordination compounds of 3,5-dinitrotyrosine and their
nonlinear optical properties. Chem Eur J 11:988–994
26. Anthony SP, Radhakrishnan TP (2004) Helical and network coordination polymers based
on a novel C2-symmetric ligand: SHG enhancement through specific metal coordination.
Chem Commun 1058–1059
27. Anthony SP, Radhakrishnan TP (2004) Coordination polymers of Cu(I) with a chiral push-pull
ligand: Hierarchical network structures and second harmonic generation. Cryst Growth Des 4:
1223–1227
28. Evans OR, Xiong RG, Wang Z, Wong GK, Lin W (1999) Crystal engineering of acentric
diamondoid metal–organic coordination networks. Angew Chem Int Ed 38:536–538
29. Evans OR, Lin W (2001) Crystal engineering of nonlinear optical materials based on
interpenetrated diamondoid coordination networks. Chem Mater 13:2705–2712
30. Lin W, Ma L, Evans OR (2000) NLO-active zinc(II) and cadmium(II) coordination networks
with 8-fold diamondoid structures. Chem Commun 2263–2264
31. Fu DW, Zhang W, Xiong RG (2008) The first metal–organic framework (MOF) of
imazethapyr and its SHG, piezoelectric and ferroelectric properties. Dalton Trans 3946–3948
32. Ledoux I, Zyss J, Siegel JS, Brienne J, Lehn JM (1990) Second-harmonic generation from
non-dipolar non-centrosymmetric aromatic charge-transfer molecules. Chem Phys Lett 172:
440–444
33. Lin W, Wang Z, Ma L (1999) A novel octupolar metal–organic NLO mateiral based on a chiral
2D coordination network. J Am Chem Soc 121:11249–11250
34. Liu Y, Li G, Li X, Cui Y (2007) Cation-dependent nonlinear optical behavior in an octupolar
3D anionic metal–organic open framework. Angew Chem Int Ed 46:6301–6304
35. Lin W, Evans OR, Xiong RG, Wang Z (1998) Supramolecular engineering of chiral and
acentric 2D networks. Synthesis, structures, and second-order nonlinear optical properties
of bis(nicotinato)zinc and bis{3-[2-(4-pyridyl)ethenyl]benzoato} · cadmium. J Am Chem Soc
120:13272–13273
36. Evans OR, Lin W (2001) Rational design of nonlinear optical materials based on 2D coordi-
nation networks. Chem Mater 13:3009–3017
37. He YH, Lan YZ, Zhan CH, Feng YL, Su H (2009) A stable second-order NLO and luminescent
Cd(II) complex based on 6-hydroxynicotinic acid. Inorg Chim Acta 362:1952–1956
38. Shi JM, Xu W, Liu QY, Liu FL, Huang ZL, Lei H, Yu WT, Fang Q (2002) Polynitrile-
bridged two-dimensional crystal: Eu(III) complex with strong fluorescence emission and
NLO property. Chem Commun 756–757
39. Han L, Hong M, Wang R, Luo J, Lin Z, Yuan D (2003) A novel nonlinear optically active
tubular coordination network based on two distinct homo-chiral helices. Chem Commun
2580–2581
40. Hu S, Zou HH, Zeng MH, Wang QX, Liang H (2008) Molecular packing variation of
crimpled 2D layers and 3D uncommon 65 · 8 topology: effect of ligand on the construction
of metal-quinoline-6-carboxylate polymers. Cryst Growth Des 8:2346–2351
Metal–Organic Frameworks for Second-Order Nonlinear Optics 165

41. Zhou WW, Chen JT, Xu G, Wang MS, Zou JP, Long XF, Wang GJ, Guo GC, Huang JS
(2008) Nonlinear optical and ferroelectric properties of a 3-D Cd(II) triazolate complex with a
novel (63)2(610 · 85) topology. Chem Commun 2762–2764
42. Yang H, Sang RL, Xu X, Xu L (2013) An unprecedented 3-D SHG MOF material of silver(I)
induced by chiral triple helices. Chem Commun 49:2909–2911
43. Yu J, Cui Y, Wu C, Yang Y, Wang Z, O’Keeffe M, Chen B, Qian G (2012) Second-order
nonlinear optical activity induced by ordered dipolar chromophores confined in the pores of
an anionic metal–organic framework. Angew Chem Int Ed 51:10542–10545
Struct Bond (2014) 157: 167–186
DOI: 10.1007/430_2013_106
# Springer-Verlag Berlin Heidelberg 2013
Published online: 19 June 2013

Host–Guest Metal–Organic Frameworks


for Photonics

Kenji Hirai, Paolo Falcaro, Susumu Kitagawa, and Shuhei Furukawa

Abstract The assembly of metal ions and organic linkers gives the highly regulated
framework scaffolds, the so-called metal–organic framework (MOFs) or porous
coordination polymers (PCPs). MOFs offer fascinating platforms in which light
emitting components can be rationally incorporated. A variety of metal ions and
organic linkers can be used to fabricate the MOF materials with a wide range of
emissive properties. Besides their inherent luminescent properties, the permanent
porosity of MOFs enables to accommodate guest species therein. The accommo-
dation of guests in the pores results in the shift of emission wavelength, the change
of emission intensity or even the generation of new emission bands. Therefore, the
luminescent MOFs can be potentially exploited as a chemical sensor for small
molecules or ions. In this chapter, we present a variety of luminescent properties
derived from the guest accommodation in MOFs, and we discuss potential
applications of luminescent MOFs as sensing materials.

K. Hirai
Department of Synthetic Chemistry and Biological Chemistry, Graduate School
of Engineering, Kyoto University, Katsura, Nishikyo-ku, Kyoto 615-8510, Japan
P. Falcaro
Division of Materials Science and Engineering, CSIRO, Private Bag 33, Clayton South MDC,
VIC 3169, Australia
S. Kitagawa
Department of Synthetic Chemistry and Biological Chemistry, Graduate School
of Engineering, Kyoto University, Katsura, Nishikyo-ku, Kyoto 615-8510, Japan
Institute for Integrated Cell-Material Sciences, Kyoto University, Yoshida, Sakyo-ku,
Kyoto 606-8501, Japan
S. Furukawa (*)
Institute for Integrated Cell-Material Sciences, Kyoto University, Yoshida, Sakyo-ku,
Kyoto 606-8501, Japan
e-mail: shuhei.furukawa@icems.kyoto-u.ac.jp
168 K. Hirai et al.

Keywords Charge transfer  Chemical sensor  Energy transfer  Metal–organic


frameworks  Porous coordination polymers  Structural transformation

Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
2 Guest-Directed Design of Luminescent MOFs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
3 Guest-Induced Luminescence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
3.1 Charge Transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
3.2 Energy Transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
3.3 Structural Change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
4 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
4.1 Chemical Sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
4.2 Sensing of Ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
4.3 Sensing of Small Molecules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
4.4 Sensing of Gas Molecules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184

Abbreviations

1,2,4-BTC Benzene-1,2,4-tricarboxylate
2,4-DNP 2,4-Dinitrophenol
ad Adeninate
adc 9,10-Anthracenedicarboxylate
bdc 1,4-Benzendicarboxylate
bpdc Biphenyldicarboxylate
bpdc 4,40 -Biphenyldicarboxylate
BPT Biphenyl-3,40 ,5-tricarboxylate
btc 1,2,4,5-Benzenetetracarboxylate
CT Charge transfer
dabco 1,4-Diazabicyclo[2.2.2]octane
DMA N, N-dimethylaniline
DMF N, N-dimethylformamide
DMPT N, N-dimethyl-p-toluidine
dpndi N, N0 -di(4-pyridyl)-1,4,5,8-naphthalenetetracarboxydiimide
DSB Distrylbenzene
DXP N, N-bis(2,6-dimethylphenyl)-3,4:9,10-perylene tetracarboxylic
diimide
FIrpic (2-Carboxypyridyl)bis(3,5-difluoro-2-(2-pyridyl)phenyl)iridium(III)
H2dtoa Dithiooxamide
IDC Imidazole-4,5-dicarboxylate
JAST Jungle-gym analogue structure
MA N-methylaniline
MIL Matériaux de l’Institut Lavoisier
MOF Metal–organic framework
Host–Guest Metal–Organic Frameworks for Photonics 169

MT 5-Methyl-1H-tetrazole
NDC 2,6-Napthalenedicarboxylate
NP 4-Nitrophenolbdc
OBA 4,40 -Oxybis(benzoate)
ox Oxalate
PCA 4-Pyridinecaboxylate
PCP Porous coordination polymer
PDA Pyridine-2,6-dicarboxylic acid
pdc Pyridine-3,5-dicarboxylate
ppy 2-Phenylpyridine
QD Quantum dot
Rh6G Rhodamine 6G
TBAPy 1,3,6,8-Tetrakis(benzoic acid)pyrene
TNP 2,4,6-Trinitrophenol
UMCM University of Michigan Crystalline Material

1 Introduction

Metal–organic frameworks (MOFs) or porous coordination polymers (PCPs) are an


intriguing class of crystalline materials owing to a wide variety of potential
applications such as adsorption, separation, and catalysis. To date, their excellent
properties related to gas separation and storage have been intensively studied;
however, the research on MOFs is currently expanding towards a new application
wherein their porous and physical properties are synergistically linked. In particu-
lar, MOFs offer fascinating platforms for the development of solid-state lumines-
cent materials, because the light-emissive building blocks can be rationally
incorporated into the framework scaffolds. The introduction of a variety of
π-conjugated ligands and luminescent metal ions allows for the design of a wide
range of emissive phenomena. Besides the inherent luminescent properties of
MOFs, the capability to accommodate guest species in the pores leads to lumines-
cent features that have not yet been observed in other solid-state materials [1, 2].
The luminescent properties of MOFs would take an advantage in a number of
different ways of the ability to encapsulate guest species within the framework. For
instance, even non-luminescent MOFs can be additionally endowed with the
luminescent properties by encapsulating luminophores [3]. By designing the size,
shape, and chemical functionality of pores, luminophores such as dye molecules,
lanthanide ions, and quantum dots can be precisely encapsulated within the porous
framework; this fabrication method provides an efficient protocol for the introduc-
tion of desired luminescent properties into the MOFs. In addition, the luminescent
properties of the photoactive sites of MOFs are strongly affected by surrounding
guest species. The accommodation of guest species usually results in the shift of
emission wavelength or the change of emission intensity. In other case, a new
emission band can be observed as a consequence of excimer or exciplex formation.
170 K. Hirai et al.

Fig. 1 Host–guest chemistry in luminescent MOFs

The mechanisms imparting luminescent functionalities are broadly classified


into three categories: charge transfer (CT), energy transfer, and structural change.
1. The charge transfer between the organic linkers and guest species leads to the
formation of exciplex state or results in luminescence quenching. In particular,
guest species with nitro or thiol groups deprive the luminescent properties of
MOFs and quench their luminescence.
2. The energy transfer between host frameworks and guest species provides a new
pathway across the excitation and emission processes, which results in the shift
of emission wavelength, luminescence quenching or antenna effect.
3. The guest accommodation causes the structural transformation around the
photoactive centers of MOFs, which induces a difference in the luminescent
intensity. Furthermore, the guest species included in the pores often restrict the
structural vibration of the host frameworks and reduce the quenching effect on
the luminescence.
These characteristic features allow the exploitation of luminescent MOFs as
chemical sensors. The rational design of MOFs for specific sensing applications
and the investigation of their luminescent response is an emerging research field.
The purpose of this chapter is to present an overview of the recent development on
host–guest chemistry for the preparation of luminescent MOFs (Fig. 1).
Host–Guest Metal–Organic Frameworks for Photonics 171

2 Guest-Directed Design of Luminescent MOFs

The unique host–guest chemistry of MOFs can be utilized to implement additional


properties by loading the cavities with functional molecules or metal nanoparticles.
The incorporation of luminescent guest species such as lanthanide ions and fluores-
cent dyes into the pores of MOFs is one strategy to fabricate the hybrid luminescent
materials. The large cavities that accommodate dye molecules can be synthesized by a
judicious choice of the metal ions and organic linkers. Qiu et al. have successfully
synthesized a new MOF with the large one-dimensional channels of 24.5  27.9 Å
([Cd3(bpdc)3(DMF)] · 5DMF · 18H2O; bpdc ¼ 4,40 -biphenyldicarboxylate, DMF ¼
N,N-dimethylformamide). The large pores of MOF accommodated rhodamine
6G (Rh6G) dye molecules and formed a [Cd3(bpdc)3(DMF)] · Rh6G complex. The
emission spectra of the [Cd3(bpdc)3(DMF)] · Rh6G complex were recorded at
different temperatures (298–77 K). The peak position of [Cd3(bpdc)3(DMF)] · Rh6G
complex at 563 nm remained unchanged with decreasing the temperature, while
the intensity of the signal was enhanced linearly; these features reveal that
[Cd3(bpdc)3(DMF)] · Rh6G is a good candidate for applications in temperature-
sensing devices [4]. Not only the dye molecules but also the coordination complex
can be incorporated in the large pores of MOFs. Fischer et al. reported the gas-phase
loading highly emissive perylene derivative (DXP) or an iridium complex (FIrpic),
into various types of MOFs (MOF-5, MOF-177, UMCM-1, and MIL-53(Al))
(DXP ¼ N,N-bis(2,6-dimethylphenyl)-3,4:9,10-perylene tetracarboxylic diimide,
FIrpic ¼ (2-carboxypyridyl)bis(3,5-difluoro-2-(2-pyridyl)phenyl)iridium(III)) [5].
The resulting host–guest hybrid materials showed strong emission endowed by the
guest species.
The approach based on lanthanide(III)-dopant is a successful strategy to impart
luminescent properties into MOFs. Luo et al. have synthesized an MOF of
{[NH4]2[Zn(btc)] · 6H2O}n (btc ¼ 1,2,4,5-benzenetetracarboxylate) [6]. The anionic
microporous framework accommodates [NH4]+ as a counter cations. The lanthanide
ions can be introduced in the micropores by the replacement with the counter cation
of [NH4]+. The MOF was immersed in the solution of EuCl3 or Tb(ClO4)3 and
successfully accommodated Eu3+ or Tb3+ in the pores (denoted as Eu@MOF and
Tb@MOF). The lanthanide-doped MOFs were used as cation sensors. Such
lanthanide-doped MOFs were subsequently soaked in the solution of MClx and
some of the lanthanide ions were exchanged with the second metal ions (M ¼ Na+,
K+, Zn2+, Ni2+, Mn2+, Co2+, Cu2+). The luminescent intensity of the metal-ion-infused
M-Eu@MOF or M-Tb@MOF was highly dependent on the nature of metal ion: Na+,
K+, Zn2+ ions showed a negligible effect on the luminescence intensity, whereas
others showed a range of quenching effects on the luminescence intensity. In particu-
lar, Cu2+ and Co2+ ions afforded a significant quenching effect on the lanthanide-
doped MOF. This study indicates that the lanthanide-doped MOFs such as Eu@MOF
or Tb@MOF are suitable for highly selective, sensitive, and low-detection-limit
172 K. Hirai et al.

Fig. 2 Bio-MOF encapsulation and sensitization of lanthanide cations. (Reprinted from [7].
Copyright 2013 American Chemical Society)

luminescent sensors of aqueous Cu2+ or Co2+ ions. Petoud and Rosi et al. also
reported a series of lanthanide-doped MOFs via cation exchange process [7]. The
as-synthesized {[Zn8(ad)4(bpdc)6O] · 2Me2NH2 · 8DMF · 11H2O}n (denoted as
bio-MOF) accommodated [Me2NH2]+ as a counter cations (ad ¼ adeninate, bpdc ¼
biphenyldicarboxylate). The lanthanide ions can be loaded into the pores of bio-MOF
by the replacement of [Me2NH2]+ with specific cations. The bio-MOF was soaked in
the solution of Tb(NO3)3, Sm(NO3)3, Eu(NO3)3, and the lanthanide ions were suc-
cessfully encapsulated. When excited at 365 nm, the doped MOFs emitted their
distinctive colors (Eu3+, red; Tb3+, green; Sm3+, orange-pink), which were detectable
with the naked eye (Fig. 2). Despite the strong quenching effect of water molecules,
Host–Guest Metal–Organic Frameworks for Photonics 173

Fig. 3 Schematic of the preparation of the QDs@MOF-5 using the ceramic seeds. The seeds
(α-hopeite microparticles) are grown within a colloidal solution of QDs. (a–c) During the QDs are
embedded during the seed growth. (d) Subsequently the seeds are washed to remove the excess of
QDs and tested with a UV lamp (365 nm) to check the presence of the luminescent particles (e).
(f) The QD doped seeds are used for the heterogeneous nucleation of MOF-5; the luminescent
QDs@MOFs are then washed and used as molecular sieve sensor for small thiol molecules.
(g) The different effect on the luminescent properties using two different thiols (large thiol ¼
n-isopropyl acrylamide/acrylic acid/t-butyl acrylamide mercaptane; small thiol ¼ ethanethiol)
along the time is presented

the characteristic luminescence of lanthanide ions was able to be observed even in


aqueous environments; this experimental evidence indicates that the bio-MOF frame-
work sufficiently protects the lanthanide ions.
A different method was proposed by Falcaro et al. [8]; the proposed protocol
take advantage of ceramic micro-particle (α-hopetite) used as seeds for the nucle-
ation of MOFs. The α-hopeite microparticles were grown in a colloidal solution of
functional nanoparticles originating doped seeds. In particular a class of efficient
emitting nanoparticles (quantum dots, QDs) have been embedded within the
α-hopeite microparticles; subsequently, the doped-ceramic particles were used for
the nucleation of MOF-5 crystals. With this method, the fabrication of hybrid
materials of QD@MOF-5 with a core-shell structure was obtained (Fig. 3). The
combination of functional materials was used as a molecular sieve sensor. In fact,
while in a pure QDs colloidal solution the presence of any thiols affects the QDs’
luminescent properties (strong luminescent emission decay towards a quenching
mechanism). With the proposed core-shell system, only small thiols were able to
diffuse within the MOF-5 frameworks quenching the QDs luminescence.
174 K. Hirai et al.

3 Guest-Induced Luminescence

3.1 Charge Transfer

Since guest species are located nearby the organic linkers of MOFs, a new emission
is often generated by the formation of an excited-state complex between host
frameworks and guest species. The infiltration of proper guest within the frame-
work and the subsequent interaction between guest species and organic linkers
leads to the formation of an excited-state complex, the so-called exciplex. Wagner
and Zaworotko et al. have reported host–guest exciplex emission of MOFs [9]. The
MOF of {[Ni(bpy)2(NO3)2] · 2pyrene}n (bpy ¼ 4,40 -bipyridine) represents a rare
example of a coordination polymer host network with an intrinsic polarity sensitive
fluorophore of pyrene as a guest molecule. The guest molecule of pyrene was
located within the framework in an ideal geometry for the formation of exciplex
between bpy and pyrene. A strong fluorescence band was observed in the pyrene
excimer region, suggesting that the pyrene molecules interact with bpy molecules.
MOFs are superior platforms to host π-conjugated organic species; taking advan-
tage of this property, a photo-excited complex with guest species can be easily
prepared. The π-conjugated molecules play an important role in facilitating charge
or electron transfer processes. Kitagawa et al. synthesized MOFs with a π-conjugated
organic linkers, ([Zn2(adc)2(dabco)]n, denoted as JAST-6), to achieve a charge
transfer (CT) interaction in the excited state between a host framework and guest
molecules (adc ¼ 9,10-anthracenedicarboxylate, dabco ¼ 1,4-diazabicyclo[2.2.2]
octane) [10]. The anthracene derivatives are suitable molecular units for MOFs
because of the unique π–π* electron transitions that provide photoluminescent
properties. The CT interactions between anthracene and N,N-dimethylaniline
(DMA) at the excited state are well known and the case was applied to the MOF
of [Zn2(adc)2(dabco)]n. The solid-state emission spectrum of the empty state of
[Zn2(adc)2(dabco)]n displayed an emission maximum at 415 nm, with a vibrational
band at 436 nm, which can be assigned to the emission from the monomeric
anthracene. After accommodating DMA in the pores, efficient quenching occurred
in the excited-state fluorescence of monomeric anthracene and a new broad emission
band in the 400–700 nm range was detected. The emission with large Stokes shift
represents a photo-induced CT complex, an exciplex between the host anthracene
unit (electron accepter) and the guest DMA molecules (electron donor). The wave-
length of exciplex emission strongly depends on the ionization potential and the
electron affinity of the donor and acceptor molecules. In fact, [Zn2(adc)2(dabco)]n
adsorbed N-methylaniline (MA) or N,N-dimethyl-p-toluidine (DMPT) providing
blue-shifted or red-shifted exciplex emissions (Fig. 4).
In general, strong electron withdrawing groups of the guest species often result in
luminescence quenching. Chen et al. reported quenching effect of nitrobenzene on
the MOF of [Zn4(OH)2(1,2,4-BTC)2]n (1,2,4-BTC ¼ benzene-1,2,4-tricarboxylate).
The luminescence of the MOF was quenched by nitrobenzene molecules inside the
pores. The quenching effect is attributed to the charge transfer electron transitions
Host–Guest Metal–Organic Frameworks for Photonics 175

Fig. 4 A snapshot of JAST-6, JAST-6MA, JAST-6DMA, and JAST-6DMPT. The samples


were subjected to UV irradiation. (Reprinted from [10]. Copyright 2013 Royal Society of
Chemistry)

from the benzene ring of BTC ligands to nitrobenzene due to the electron-deficient
property of nitrobenzene and the π–π interactions between the framework and
nitrobenzene molecules [11].

3.2 Energy Transfer

The energy transfer between the host frameworks and guest species often results
in the shift of emission wavelength, luminescence quenching, or antenna effect.
Uvdal et al. have synthesized an MOF comprising of π-conjugated dicarboxylate
linkers (L) and a series of lanthanide metal ions (Gd(III), Eu(III) and Yb(III))
[12, 13]. Because the local environment around Ln(III) will be negatively charged
due to the binding of the acetate anion to the lanthanide ion, the positively charged
dye molecule of trans-4-styryl-1-methylpyridiniumiodide (D) was encapsulated in
the MOF. The D-loaded MOF exhibited an efficient light-harvesting effect. The
overlap between the emission spectrum of L itself (donor) and the absorption of
D (acceptor) favors the Förster-type energy transfer between them [14]. The emis-
sion from the D in the MOF was strongly enhanced compared with direct excitation
of the D dye. Furthermore, the delay in fluorescent decay curves was observed in
time-resolved fluorescence spectroscopy. Therefore, D was not directly excited by
the laser pulse, suggesting the energy transfer from L to D. The accommodation of
dye molecules in the frameworks based on such π-conjugated linkers allows for a
charge transfer couple between the organic linkers and guest species, thus leading
to the highly efficient light harvesting system.
176 K. Hirai et al.

Qian et al. reported an MOF of {[Tb(BTC)] · (DMF) · (H2O)}n (MOF-76 · guest)


for the sensing of DMF and acetone [15]. The luminescent intensity of partially
desolvated MOF in n-propanol increased and decreased with the additions of DMF
and acetone solvents, respectively; remarkably, the luminescent intensity did not
change in the cases of other solvents. Both absorption and luminescent spectra
showed that the energy absorbed by BTC is transferred to acetone molecules,
resulting in decreasing the efficiency of intersystem crossing S1 ! T1, where S1
and T1 are the singlet state and triplet state of BTC, respectively. In comparison with
the absorption peak of BTC, the absorption band of DMF was located higher
frequencies; this leads to an efficient energy transfer from DMF to BTC and results
in larger efficiency of intersystem crossing S1 ! T1.

3.3 Structural Change

In general, small structural differences around the luminescent centers give signifi-
cant change in the emission properties. The guest accommodation often results in the
structural transformation of MOFs, thereby leading to the shift of emission wave-
length and/or intensity changes. Duan et al. reported an MOF of {[Cu6L6] · (H2O)
(DMSO)}n, which exhibited highly selective absorption for aromatic molecules in
water and different fluorescent quenching effect for toluene, nitrobenzene, aniline,
and o-, p-, and m-dimethylbenzene (HL ¼ 5,6-diphenyl-1,2,4-triazine-3-thiol) [16].
The luminescent spectrum of the desolvated MOF exhibited a cluster-centered band
at 660 nm because of the luminescent properties of Cu6S6 clusters. Such critical
photophysical properties were highly dependent on the Cu–Cu distance within the
cluster; therefore, the emission intensities could be modified by changing the intra-
cluster Cu–Cu distance by the inclusion of interacting guests. The accommodation
of aromatic molecules within the MOF led to the system exhibiting no obvious
luminescence, indicating that aromatic molecules quenched the luminescence of
the cluster. The monosubstituted benzene derivatives, toluene, nitrobenzene, aniline
in the pores of MOF reduced the luminescent intensity to a 25% of its original
amplitude. Thus, the characteristics of substituent groups are not main factors
for quenching the luminescence. Furthermore, o-, p-, and m-dimethylbenzene gave
different luminescent intensities. The molecular structures of monosubstituted
benzene and o-dimethylbenzene are suitable for the formation of π–π stacking
between the host frameworks and guest species. As a result, the Cu–Cu distance
was enlarged and the cluster-centered luminescence of Cu6L6 was quenched
effectively.
The luminescence of lanthanide coordination compounds is generated through a
three steps antenna effect. Firstly, light is absorbed by the organic ligands;
secondly, the energy is transferred to LnIII; finally, the LnIII emits the luminescence.
The vibration of organic linkers reduces the efficiency of the energy transfer from
the ligand to LnIII. Lu et al. reported the guest-induced intensity enhancement
of lanthanide-based MOF. The lanthanide-based MOF of K5[Tb5(IDC)4(ox)4]
Host–Guest Metal–Organic Frameworks for Photonics 177

accommodated metal cations by the exchange of K+ ion (IDC ¼ imidazole-4,5-


dicarboxylate, ox ¼ oxalate) [17]. The intensity of the 5D4 ! 7 F5 emission of
Tb3+ is significantly enhanced by the addition of Ca2+ ions. Furthermore, the
fluorescence lifetime increased from 158.90 to 287.48 μs as a consequence of the
addition of Ca2+. The proper amount of Ca2+ bonded to ox2 and the interactions
between Ca2+ and ox2 suppresses the vibration of ox2, thus increasing the
efficiency of the energy transfer from the ligand to TbIII. Chen et al. recently
demonstrated the sensing property of MOF for Cu2+ ion in the simulated physio-
logical aqueous solution [18].

4 Applications

4.1 Chemical Sensors

Molecular recognition with the ability to distinguish among small structural


differences of molecules is a critical and challenging task for a variety of applica-
tive fields such as biotechnology, medicine, industrial processing, and environmen-
tal monitoring. In particular, the sequential system of recognition-transduction to
detect these differences is a key to fabricate sensor materials for explosive, harmful,
or pollutant small molecules. Most chemical sensors have been developed by
following this key principle: recognition of specific molecules and transduction to
convert the recognition event into a detectable signal. Luminescent MOFs are one
of the promising materials to meet the requirement of this key principle. This is
because the event of guest accommodation gives the differences in the luminescent
emission: the shift of emission wavelength, the change of emission intensity, or the
emergence of new emission as a result of exciplex formation. Therefore, MOFs can
be used as chemical sensing platform to identify the specific guest molecules.
A remarkable advantage of MOFs as chemical sensors relies on the designability
of luminescent property by a proper choice of components. The combination of the
inherent porosity and tunable luminescent properties of MOFs provides a fascinating
opportunity for the fabrication of an efficient class of sensing materials. Herein, we
present the sensing ability of MOFs for metal ions, organic molecules, and gas
molecules.

4.2 Sensing of Ions

Distinguishing and detecting metal ions is the key challenges in particular for envi-
ronmental monitoring and biotechnology. One way to detect specific metal ions can
be implemented by the introduction of interactive sites for the target ions in MOFs.
178 K. Hirai et al.

Liu et al. first reported the cation sensing by the lanthanide-based MOF of {Na[EuL
(H2O)4] · 2H2O}n (L ¼ 1,4,8,11-tetraazacyclotetradecane-1,4,8,11-tetrapropionate)
[19]. The effects of a variety of cations (Cu2+, Ag+, Zn2+, Cd2+, and Hg2+) on the
fluorescence spectrum of the MOF were investigated. The cations can diffuse within
the framework and occupy the empty coordination site in the azacycle of the
ligand (L). The cations of Cu2+, Zn2+, Cd2+, and Hg2+ quenched the fluorescence
of the Eu-based MOF. In contrast, Ag+ gave two effects: the enhancement of the
emission and the change of the emission spectrum of Eu3+ from multiple peaks to a
single peak. The emission of 5D0 ! 7F2 transition was enhanced, but other
emissions of 5D0 ! 7FJ transition (J ¼ 1, 4) was diminished. This is probably
because the accommodation of Ag+ in azacycle moiety of the ligand enhanced the
rigidity and modified the paramagnetic spin state of the MOF. Chen et al. also
synthesized the MOF of {[Eu(pdc)1.5] · (DMF)}n, comprising of the interaction
site for cations (pdc ¼ pyridine-3,5-dicarboxylate) [20]. The metal ions were cap-
tured by the pyridine moiety of pdc. While alkali and alkaline earth metal ions have a
negligible effect on the luminescence intensity, Co2+ and Cu2+ significantly reduced
the luminescence intensity. As an explanation for the quenching of the luminescence,
Chen et al. suggest that the binding of the pyridyl nitrogen atom to Cu2+ or Co2+
reduces the efficiency of antenna effect of pdc to magnify the f–f transitions of Eu3+.
It is known that the luminescent intensity of lanthanide ions strongly depends on
the efficiency of energy transfer from the ligand to lanthanide ions. Cheng et al.
reported the ion sensing by the lanthanide-based MOF of [Eu(PDA)3Mn1.5(H2O)3]n
and [Tb(PDA)3Mn1.5(H2O)3]n (PDA ¼ pyridine-2,6-dicarboxylic acid) [21]. The
luminescent intensity of the MOFs was strongly enhanced by the addition of Zn2+
ions; however, no change was detected by the addition of Ca2+ and Mg2+ ions. While
the addition of Mn2+ ions decreased the luminescent intensity, Fe2+, Co2+, and Ni2+
quenched the luminescence. The authors speculated that the addition of metal ions
changed the photo-induced electron transfer process from the ligand to the Eu3+
or Tb3+ ions, thus resulting in the variety of effects of ions on the luminescent
intensity. Cheng et al. also reported an Ln-based MOF of [Eu(PDA)3Fe1.5(H2O)3]n
for metal ions sensing. Unlike the previous report, the presence of Zn2+ decreased
the luminescent intensity of the 5D0 ! 7 F2 transition while Mg2+ increased the
emission intensity [22].
A number of studies on the anion sensing by luminescent MOFs have been also
reported. Wong et al. synthesized an MOF of [Tb(Mucicate)1.5 · 3(H2O)2]n, which
serves as a luminescent receptor for selective anion monitoring [23]. The lumines-
cent intensity of the Tb3+ was changed by immersing MOFs in an aqueous solution
of sodium salts. Anions such as I, CO32, CN increased the luminescent inten-
sity, while SO42 and PO43 did not show any effect. The response of the channels
to anions is probably related to the numerous OH groups in the mucic acid; the
oxydril groups would possibly strongly interact with the anions through hydrogen
bonding. The strong interaction between OH group and anions suppresses the
quenching effect of OH group on the luminescence of lanthanide. Chen et al.
demonstrated another method to detect anions by the luminescent change of
lanthanide-based MOF [24]. MOF-76b of {[Tb(BTC)3] · (methanol)}n showed
Host–Guest Metal–Organic Frameworks for Photonics 179

Fig. 5 (a) Single crystal structure of MOF-76b activated in methanol containing NaF with the
model of fluoride at the center of the channel involving its hydrogen-bonding interaction with
terminal methanol molecules (methanol oxygen, purple; the methyl group from methanol is
omitted for clarity). (b) Excitation (dotted) and PL spectra (solid) of MOF-76b solid activated in
different concentrations of NaF methanol solution (excited and monitored at 353 and 548 nm,
respectively). (Reprinted from [24]. Copyright 2013 American Chemical Society)

the luminescent enhancement upon the addition of Br, Cl, F, SO42, and CO32
in methanol (BTC ¼ 1,3,5-benzenetricarboxylate). These anions weakened the
vibrational quenching effect of OH group due to the interaction with hydrogen
atom of methanol, thus enhances the luminescence intensity. In particular, F ion
strongly interacted with the hydrogen atom, showing strongest significant enhance-
ment of the luminescent intensity (Fig. 5).
Deng et al. demonstrated that transition metal ion-based MOFs can be used for
the detection of anions [25]. The MOF of [Cd(μ2-Cl)(μ4-5MT)]n exhibits an emis-
sion at around 370 nm owing to the intra-ligand transitions (MT ¼ 5-methyl-1H-
tetrazole). The decrease in the emission intensity caused by the addition of NO2 is
attributed to the energy transfer mechanism between the methyl groups and the
nitrite ions.

4.3 Sensing of Small Molecules

The chemical functionality and structure of guest molecules significantly influence


the luminescent emissions of MOFs. Rosseinsky et al. reported a pyrene-based
MOF of [In2(OH)2(TBAPy)]n (TBAPy ¼ 1,3,6,8-Tetrakis(benzoic acid)pyrene),
which showed a wide range of solvent-dependent luminescence [26]. The emission
band for the framework was shifted to lower energy by the desolvation. The
emission intensity and the lifetime were dependent on the amount and the chemical
nature of the species hosted in the pores. Polar DMF produced a high degree of
structural order and separates the linkers to minimize self-quenching phenomena,
resulting in a fine emission profile. Thanks to the hydrogen-bonding of the polar
180 K. Hirai et al.

water guest and carboxylate groups, a framework structure was distorted, resulting
in distinct emission properties (e.g., different properties are found using DMF as a
solvent). In particular, the structural distortions induced by the interaction with
polar guests strongly affected the lifetime of the emission. Different nonpolar
molecules such as dioxane and toluene guests induced a limited distortion of the
framework structure; however, as confirmed by diffraction analyses, the specific
π-stacking interaction between the toluene and the pyrene moiety results in the
reduced emission intensity.
The MOFs based on luminescent open metal sites exhibit attractive sensing
capability for solvent molecules. Chen et al. reported the luminescent intensity
change of an MOF induced solvent molecules coordinating an open metal site [27].
The MOF of [Eu(BTC)(H2O)3]n possesses an open metal site of Eu3+ (BTC ¼
benzenetricarboxylate), which acts as a recognition site for guest molecules. While
open Eu3+ metal site coordinated by DMF molecules showed an increased lumines-
cent intensity related to the 5D0 ! 7F2 transition, the coordination operated by
acetone to the metal site significantly decreased the intensity. Chen et al. demonstrated
a further example of the luminescent intensity change triggered by the coordination
of guest molecules to open metal sites [28]. An MOF based on {[Yb(BPT)(H2O)] ·
(DMF)1.5(H2O)1.25}n, (BPT ¼ biphenyl-3,40 ,5-tricarboxylate) shows an open metal
site once guest molecules and coordinated water molecule are removed. The
desolvated MOF of [Yb(BPT)]n exhibited emission from the 2F5/2 ! 2F7/2 transition
of the Yb3+ ion. As reported in the previous work [27], DMF and acetone most
significantly enhanced and quenched the luminescence, respectively.
Hsu et al. demonstrated the sensing property of an Na[Tb(OBA)2]n (OBA ¼ 4,40 -
oxybis(benzoate)) MOF [29]; the proposed MOF showed the ability to act as a
stimuli responsive luminescent material towards a variety of solvents. The lumines-
cence measurements of the MOF displayed that the 5D4 ! 7F4 emissions were the
strongest intensities in BuOH and EtOH suspensions, but these intensities were
significantly reduced in MeOH and H2O suspensions. The small molecules of H2O
and MeOH get closer to the terbium ions of the framework that quenches the
luminescence intensities effectively; on the other hand, the alkyl sterical hindrance
of EtOH and BuOH potentially protects the terbium ions from quenching by OH
oscillators.
In general, OH oscillators in the water molecules act as quenching species for the
lanthanide luminescence. However, Wang and Gao et al. utilized water molecules to
enhance the luminescent intensity of lanthanide-based MOFs [30]. The as-synthesized
MOFs of {[Ln2(fumarate)2(oxalate)(H2O)4] · 4H2O}n (Ln ¼ Eu, Tb) accommodated
water molecules as guest species and showed luminescence derived from the lantha-
nide ions. The dehydration of the MOFs significantly reduced the crystallinity with a
partial loss of the luminescent properties; after a rehydration process the lumines-
cence was recovered. This is probably due to the softened environment of lanthanide
ion in the dehydrated amorphous phases and the recovery of the stiffness after
rehydration. A similar sensing behavior has been discovered by Li et al. using needle
Host–Guest Metal–Organic Frameworks for Photonics 181

shape crystals of {[Eu2L3(H2O)4] · 3DMF}n, (L ¼ 20 ,50 -bis(methoxymethyl)-


[1,10 :40 ,100 -terphenyl]-4,400 -dicarboxylate) [31]. The proposed Eu-based MOF presents
an emission spectra corresponding to the 5D0 ! 7F0, 5D0 ! 7F1, 5D0 ! 7F2, and
5
D0 ! 7F4 transitions. Remarkably, the MOF shows an eightfold enhancement of
the luminescence intensity in presence of DMF vapors. The luminescent signal
becomes weaker once the MOF is exposed to water vapors. Thus, the authors
demonstrated sensor response of the MOF to DMF and water vapors. The vapors
of DMF and water were introduced into the sensor cell in an alternating manner, and
the luminescent intensity was monitored continuously. The turn-on and turn-off
switching of luminescence was achieved within 10–20 s.
Molecules containing nitro or thiol groups often serve as efficient quencher
because of the electron withdrawing properties or the hole trapping ability, respec-
tively. Ghosh et al. reported a luminescent MOF of [Cd(NDC)0.5(PCA)]n (NDC ¼
2,6-napthalenedicarboxylate, PCA ¼ 4-pyridinecaboxylate) for the highly selec-
tive detection of the nitro explosive TNP (TNP ¼ 2,4,6-trinitrophenol) [32]. The
proposed MOF exhibited selective detection of TNP regardless the presence of other
nitro compounds in both aqueous and organic solutions. The selectivity is ascribed to
electron and energy transfer mechanisms, as well as electrostatic interactions
between TNP and the luminescent center. The order of the quenching efficiency
was found to be TNP 2,4-DNP > NP, which is in agreement with the order of
acidity of the molecules (2,4-DNP ¼ 2,4-dinitrophenol, NP ¼ 4-nitrophenol). TNP
with the highly acidic hydroxy group interact strongly with the luminescent center
and efficiently quench the luminescence due to the favorable electron and energy
transfer mechanism. Falcaro et al. reported luminescent hybrid materials by the
encapsulation of QD in MOF-5, which detect thiol molecules by the quenching
effect [8, 33]. Only thiols with the molecular size smaller than the pore were
observed to affect the luminescence of the QDs. These studies suggested that the
encapsulation of luminophores in MOF is also a successful method to fabricate
molecular sieve MOF-based sensors.
The structural flexibility of MOFs, which allows for the confinement of
molecules with maximized framework-guest interaction, is a prominent feature
that differentiates MOFs from other conventional porous solids. Although the
MOF dynamic structure holds promise for applications in storage and separation,
MOFs have shown the potential to be used as an efficient chemical sensing
platform, thanks to the framework’s ability to adapt the nanopore depending on
the analyte features. Furukawa and Kitagawa et al. demonstrated the guest-induced
luminescence of dpndi-based interpenetrated MOF of [Zn2(bdc)2(dpndi)]n (bdc ¼
1,4-benzendicarboxylate, dpndi ¼ N,N0 -di(4-pyridyl)-1,4,5,8-naphthalenetetracar-
boxydiimide) [34]. The interpenetrated framework undergoes the dynamic struc-
tural transformation to confine a class of aromatic volatile organic compounds. The
displacement of two chemically non-interconnected frameworks provides space for
trapping the guest molecules. The confinement of guest species enhances the
interaction of aromatic guest molecules with the dpndi embedded in the framework
182 K. Hirai et al.

Fig. 6 The resulting luminescence of crystal powders, suspended in each liquid after excitation at
370 nm: benzene (blue), toluene (cyan), xylene (green), anisole (yellow), iodobenzene (red)

scaffolds. This leads to the formation of complexes with a CT nature, producing


fluorescence from the exciplex or the stabilization of the triplet state of dpndi to
generate phosphorescence. Although the MOF of [Zn2(bdc)2(dpndi)]n shows a very
weak fluorescence with a very low quantum yield, the interaction between dpndi
and aromatic guest molecules generated the visible fluorescence from violet to
yellow by increasing the donor guest ability, as shown in Fig. 6. Furthermore, the
interaction between dpndi and iodobenzene led to red emission by phosphorescence
stabilized by heavy atom effect. Since the emission color depends on the chemical
substitution on the phenyl ring, the MOF can identify the aromatic molecules by
visible fluorescence.
Chen et al. established a new methodology to detect large biomolecules by using
an MOF of [Cu(H2dtoa)]n (H2dtoa ¼ dithiooxamide) [35]. In the absence of MOF,
free probe DNA with a luminophore at its 30 end has strong luminescence at an
excitation wavelength of 480 nm. The introduction of MOF bound with probe DNA
as a result of hydrophobic and π-stacking interactions between the nucleobases and
the MOF quenched the fluorescence by an energy transfer process. In the presence
of a target such as target DNA or protein the binding between the target and probe
DNA alters the conformation of probe DNA. This structural alteration releases
probe DNA from the MOF, which results in the restoration of fluorescence. Based
on this sensing platform, HIV DNA and thrombin can be detected. The system has
shown a high sensitivity, in fact the detection limits are 3 nM for the HIV DNA and
1.3 nM for the thrombin.
Host–Guest Metal–Organic Frameworks for Photonics 183

4.4 Sensing of Gas Molecules

There have been many examples of gas detection that uses luminescence quenching
by energy transfer between gases and luminescent materials. However, only a
limited number of gases, such as O2 and NO2, cause luminescence quenching.
Thus, not so many studies on gas detection have been reported in luminescent
MOFs. Lin et al. demonstrated O2 induced luminescence quenching of [Zn4(μ4-O)
(Ir(ppy)3 derivatives)2]n [36]. The MOF exhibited the luminescence from Ir(ppy)3
(ppy ¼ 2-phenylpyridine) moieties. The MLCT phosphorescence can be readily
quenched by molecules with a triplet ground state through the incorporation of O2
into the framework. Thus, the luminescence of MOF is reversibly quenched by O2.
Petoud and Rosi et al. reported lanthanide doping in bio-MOF via cation exchange
process [7]. The lanthanide-doped MOF also showed the capability of O2 sensing.
Uemura and Kitagawa et al. demonstrated an MOF-based detection of CO2
[37]. The photoactive molecules of distrylbenzene (DSB ¼ distrylbenzene) were
introduced into the MOF of [Zn2(bdc)2(dabco)]n. The framework structure of MOF
was distorted by accommodating DSB, and the structural conformation of DSB was
twisted. Because the twisting of DSB affects the π-conjugation efficiency across the
benzene rings, the MOF accommodating DSB did not show the luminescence.
However, CO2 adsorption into the extra space of pores induced the structural
transformation of the MOF from distorted framework to original tetragonal frame-
work. This structural transformation caused the conformational transformation of
DSB from the twisted arrangement to the original linear one. As a result, the CO2
adsorption triggers the luminescence emission of MOF-DSB hybrid materials.
The flexible MOF and DSB synergistically collaborated to detect CO2.

5 Conclusion

Luminescent MOFs are promising materials for chemical sensing, light-emitting


devices, and biotechnology; however, the research on luminescent MOFs is still at
an early stage. Although the host–guest-dependent luminescence mechanism
related is now under exploration and new rational design of MOFs for specific
sensing applications is becoming a raising research field, the majority of the
literature reports are limited to the presentation of the luminescent properties.
The designability of framework structures enables the manipulation of their inher-
ent luminescent properties. Furthermore, the nanoporosity of MOFs allows for
guest accommodation enabling the exploitation of guest-dependent luminescent
properties. In the near future, if the luminescent sensing properties would be
combined with the sieving function of the porous MOFs, sophisticated luminescent
porous materials would be fabricated [38]. Moreover, the detailed mechanism of
guest-dependent luminescence is going to be revealed thanks to the availability of
fluorescence facilities with the ability to measure specific features including
184 K. Hirai et al.

excitation spectra, lifetimes, or quenching rate constants. Moreover, the miniaturi-


zation of spectrometers make very promising the perspective to use portable
sensing devices based on MOFs. The close multidisciplinary collaboration among
chemists, materials scientists, photophysists, and engineers will certainly facilitate
the implementation of some promising luminescent MOFs and technologies for
practical applications in chemical sensing.

References

1. Allendorf MD, Bauer CA, Bhakta RK, Houk RJT (2009) Luminescent metal–organic
frameworks. Chem Soc Rev 38:1330–1352
2. Cui Y, Yue Y, Qian G, Chen B (2012) Luminescent functional metal–organic frameworks.
Chem Rev 112:1126–1162
3. Falcaro P, Furukawa S (2012) Doping light emitters into metal–organic frameworks. Angew
Chem Int Ed 51:8431–8433
4. Fang QR, Zhu GS, Jin Z, Ji YY, Ye JW, Xue M, Yang H, Wang Y, Qiu SL (2007) Mesoporous
metal–organic framework with rare etb topology for hydrogen storage and dye assembly.
Angew Chem Int Ed 46:6638–6642
5. Müller M, Devaux A, Yang CH, De Cola L, Fischer RA (2010) Highly emissive metal–organic
framework composites by host–guest chemistry. Photochem Photobiol Sci 9:846–853
6. Luo F, Batten SR (2010) Metal–organic framework (MOF): lanthanide(III)-doped approach
for luminescence modulation and luminescent sensing. Dalton Trans 39:4485–4488
7. An J, Shade CM, Chengelis-Czegan DA, Petoud S, Rosi NL (2011) Zinc-adeninate metal–organic
framework for aqueous encapsulation and sensitization of near-infrared and visible emitting
lanthanide cations. J Am Chem Soc 133:1220–1223
8. Falcaro P, Hill AJ, Nairn KM, Jasieniak J, Mardel JI, Bastow TJ, Mayo SC, Gimona M, Gomez D,
Whitfield HJ, Riccò R, Patelli A, Marmiroli B, Amenitsch H, Colson T, Villanova L, Buso D
(2011) A new method to position and functionalize metal–organic framework crystals. Nat
Commun 2:237
9. Wagner BD, McManus GJ, Moulton B, Zaworotko MJ (2002) Exciplex fluorescence of
{[Zn(bipy)1.5(NO3)2}] · CH3OH · 0.5pyrene}n: a coordination polymer containing intercalated
pyrene molecules (bipy ¼ 4,40 -bipyridine). Chem Commun 2176–2177
10. Tanaka D, Horike S, Kitagawa S, Ohba M, Hasegawa M, Ozawa Y, Toriumi K (2007)
Anthracene array-type porous coordination polymer with host–guest charge transfer
interactions in excited states. Chem Commun 3142–3144
11. Zhang Z, Xiang S, Rao X, Zheng Q, Fronczek FR, Qian G, Chen B (2010) A rod packing
microporous metal–organic framework with open metal sites for selective guest sorption and
sensing of nitrobenzene. Chem Commun 46:7205–7207
12. Zhang X, Ballem MA, Ahrén M, Suska A, Bergman P, Uvdal K (2010) Nanoscale Ln(III)-
carboxylate coordination polymers (Ln ¼ Gd, Eu, Yb): temperature-controlled guest encap-
sulation and light harvesting. J Am Chem Soc 132:10391–10397
13. Zhang X, Ballem MA, Hu ZJ, Bergman P, Uvdal K (2011) Nanoscale light-harvesting
metal–organic frameworks. Angew Chem Int Ed 50:5729–5733
14. Wu P, Brand L (1994) Resonance energy transfer: methods and applications. Anal Biochem
218:1–13
15. Xiao Y, Wang L, Cui Y, Chen B, Zapata F, Qian G (2009) Molecular sensing with lanthanide
luminescence in a 3D porous metal–organic framework. J Alloys Comp 484:601–604
16. Bai Y, He GJ, Zhao YG, Duan CY, Dang DB, Meng QJ (2006) Porous material for absorption
and luminescent detection of aromatic molecules in water. Chem Commun 1530–1532
Host–Guest Metal–Organic Frameworks for Photonics 185

17. Lu WG, Jiang L, Feng XL, Lu TB (2009) Three-dimensional lanthanide anionic metal–organic
frameworks with tunable luminescent properties induced by cation exchange. Inorg Chem
48:6997–6999
18. Xiao Y, Cui Y, Zheng Q, Xiang S, Qian G, Chen B (2010) A microporous luminescent
metal–organic framework for highly selective and sensitive sensing of Cu2+ in aqueous
solution. Chem Commun 46:5503–5505
19. Liu W, Jiao T, Li Y, Liu Q, Tan M, Wang H, Wang L (2004) Lanthanide coordination
polymers and their Ag+-modulated fluorescence. J Am Chem Soc 126:2280–2281
20. Chen B, Wang L, Xiao Y, Fronczek FR, Xue M, Cui Y, Qian G (2009) A luminescent
metal–organic framework with Lewis basic pyridyl sites for the sensing of metal ions.
Angew Chem Int Ed 48:508–511
21. Zhao B, Chen XY, Cheng P, Liao DZ, Yan SP, Jiang ZH (2004) Coordination polymers
containing 1D channels as selective luminescent probes. J Am Chem Soc 126:15394–15395
22. Zhao B, Chen XY, Chen Z, Shi W, Cheng P, Yan SP, Liao DZ (2009) A porous 3D
heterometal–organic framework containing both lanthanide and high-spin Fe(II) ions. Chem
Commun 3113–3115
23. Wong KL, Law GL, Yang YY, Wong WT (2006) A highly porous luminescent terbium-
organic framework for reversible anion sensing. Adv Mater 18:1051–1054
24. Chen B, Wang L, Zapata F, Qian G, Lobkovsky EB (2008) A luminescent microporous
metal–organic framework for the recognition and sensing of anions. J Am Chem Soc
130:6718–6719
25. Qiu Y, Deng H, Mou J, Yang S, Zeller M, Batten SR, Wue H, Lie J (2009) In situ tetrazole
ligand synthesis leading to a microporous cadmium-organic framework for selective ion
sensing. Chem Commun 5415–5417
26. Stylianou KC, Heck R, Chong SY, Bacsa J, Jones JTA, Khimyak YZ, Bradshaw D, Rosseinsky
MJ (2010) A guest-responsive fluorescent 3D microporous metal–organic framework derived
from a long-lifetime pyrene core. J Am Chem Soc 132:4119–4130
27. Chen B, Yang Y, Zapata F, Lin G, Qian G, Lobkovsky EB (2007) Luminescent open metal sites
within a metal–organic framework for sensing small molecules. Adv Mater 19:1693–1696
28. Guo Z, Xu H, Su S, Cai J, Dang S, Xiang S, Qian G, Zhang H, O’Keeffe M, Chen B (2011) A
robust near infrared luminescent ytterbium metal–organic framework for sensing of small
molecules. Chem Commun 47:5551–5553
29. Lin YW, Jian BR, Huang SC, Huang CH, Hsu KF (2010) Synthesis and characterization of
three ytterbium coordination polymers featuring various cationic species and a luminescence
study of a terbium analogue with open channels. Inorg Chem 49:2316–2324
30. Zhu WH, Wang AM, Gao S (2007) Two 3D porous lanthanide–fumarate–oxalate frameworks
exhibiting framework dynamics and luminescent change upon reversible de- and rehydration.
Inorg Chem 46:1337–1342
31. Li Y, Zhang S, Song D (2013) A luminescent metal–organic framework as a turn-on sensor for
DMF vapor. Angew Chem Int Ed 52:710–713
32. Nagarkar SS, Joarder B, Chaudhari AK, Mukherjee S, Ghosh SK (2013) Highly selective
detection of nitro explosives by a luminescent metal–organic framework. Angew Chem Int Ed
52:2881–2885
33. Buso D, Jasieniak J, Lay MDH, Schiavuta P, Scopece P, Laird J, Amenitsch H, Hill AJ, Falcaro P
(2012) Highly luminescent metal–organic frameworks through quantum dot doping. Small
8:80–88
34. Takashima Y, Martı́nez VM, Furukawa S, Kondo M, Shimomura S, Uehara H, Nakahama M,
Sugimoto K, Kitagawa S (2011) Molecular decoding using luminescence from an entangled
porous framework. Nat Commun 2:168
35. Zhu X, Zheng H, Wei X, Lin Z, Guo L, Qiu B, Chen G (2013) Metal–organic framework
(MOF): a novel sensing platform for biomolecules. Chem Commun 49:1276–1278
36. Xie Z, Ma L, deKrafft KE, Jin A, Lin W (2010) Porous phosphorescent coordination polymers
for oxygen sensing. J Am Chem Soc 132:922–923
186 K. Hirai et al.

37. Yanai N, Kitayama K, Hijikata Y, Sato H, Matsuda R, Kubota Y, Takata M, Mizuno M,


Uemura T, Kitagawa S (2011) Gas detection by structural variations of fluorescent guest
molecules in a flexible porous coordination polymer. Nat Mater 10:787–793
38. Hirai K, Furukawa S, Kondo M, Meilikhov M, Sakata Y, Sakata O, Kitagawa S (2012)
Targeted functionalisation of a hierarchically-structured porous coordination polymer crystal
enhances its entire function. Chem Commun 48:6472–6474
Index

A Cation sensors, 64
Acetylene, 18 Charge transfer, 167, 170, 174
Aggregation-induced emission luminescence, 46
(AIE), 40 Chelidonic acid, 118
Alkenoates, 105, 139 Chemical sensors, 27, 63, 167, 177
Anion sensors, 64 Cisplatin, 76
Anthracene tetracarboxylate, 18 Cluster-based frameworks, 5
Anthrax, 68 Co(II) imidazolate, 9
Azobenzene, cis–trans isomerization, 134 Coordination polymers (CPs), 2, 105
Correlated color temperature (CCT), 60
Crystal engineering, 105
B Cycloaddition, 105
Bacillus anthracis, 68 Cyclobutane derivatives, 105
Barium borate, 147 Cyclohexane tetracarboxylic acid, 33
Benzenedicarboxylate, 6, 33, 68, 181 Cyclopentanones, 107
Benzenetricarboxylate, 6, 34, 36, 39, 65,
174, 180
Biomedicine, 27 D
Bipyridine, 33 Diamines, 3
Bis(imidazol-1-ylmethyl)benzene, 149 Diamondoid structures, 152
1,3-Bis(3-phenylprop-2-enol-oxo-2,4-bis net, 145
(phenyl)cyclobutane), 107 Dicarboxylate-pyridylcarboxylate
Bis(4-phenyl-2-pyridine)(5,50 -di ligand, 7
(4-phenylcarboxylate)-2,20 -bipyridine)- Dicarboxylates, bioactive, 118
iridium(III) chloride, 91 Diketonates, 58
Bis(4-phenyl-2-pyridine)(5,50 -dicarboxylate)- Dimethyldinitrobutane, 71
2,20 -bipyridine)-iridium(III) Dimethylpyrazol-4-yl benzoic acid, 20
chloride, 91 Dinitriles, 3
Bis(3-(4-pyridyl)acryloyl)-hydrazine, 117 Dinitrophenol, 181
Bis(40-pyridyl)2,4-bis(phenyl)cyclobutane, 5,6-Diphenyl-1,2,4-triazine-3-thiol, 176
120 Dipicolinic acid, 68
Bis[2-(4-pyridyl)ethenyl]benzene, 132 Di(4-pyridyl)
naphthalenetetracarboxydiimide), 181
Distrylbenzene, 183
C Divinyl-benzene-1,4-dicarboxylate
Carbon dioxide, 19, 183 (DVTP), 134
reduction, 89, 93 DNA, probe, 182

187
188 Index

E guest-induced, 174
Energy transfer, 167, 175 ligand-based, 38
Ethoxysuccinato-cisplatin (ESCP), 76 Luminophors, 22
Explosives, 71

M
F Magnetic resonance imaging (MRI), 75
Framework flexibility, 15 Metal azolate frameworks (MAFs), 3
Metal ions, sensing/detecting, 177
Metal–organic frameworks (MOFs), 1, 27, 89,
G 105, 145, 167
Gas adsorption, 134 Metal-to-ligand charge transfer (MLCT),
Gas sensors, 69, 183 41, 46
Gas sorption, 125 Methane absorbance, 18
Gas storage, 16 Methyl orange (MO), 96
Grid networks, 157 5-Methyl-1H-tetrazole, 65
Guest-induced luminescence, 47 1-Methyl-2-pyrrolidone (NMP), 34
MOFs, nanoscale, 35
Molecular design, 1
H Molecular recognition, 177
1H-Imidazo[4,5-f][1,10]phenanthroline, 22 Molecular sensing, 21
HIV DNA, 182 Molecule sensors, 66
Host–guest frameworks, 2, 170 Multimodal imaging, 75
Host–guest reactions, 133
Hydrogen, 17
6-Hydroxynicotinic acid, 158 N
Nano-MOFs (NMOFs), 35
Nanoparticles, 173
I Naphthalenediimide, 69
Ideal adsorption solution theory Nitrophenol, 181
(IAST), 19 Noncentrosymmetric MOFs, 145
Imazethapyr, 154 Nonlinear optics (NLO), 145
Imidazolates, 8
Imidazoles, 12, 34, 159
Imidazole-4,5-dicarboxylate, 177 O
In–pyrenetetrabenzoic acid MOF, 67 Octupolar nonlinearities, 155
Interpenetration, 8 Octupolar symmetry, 145
Intramolecular charge transfer (ICT), 147 Oxybis(benzoate), 67, 180
Ion-and liquid-assisted grinding (ILAG), 34
Ions, sensing, 177
Iridium complexes, cyclometalated, 69 P
Isoreticular MOFs (IRMOFs), 32 Pentadiene-3-one, 118
Phosphorescence, 38
Photocatalysis, 89
L Photocatalytic degradation, 95
Ladder structures, 105, 113–124, 136–140, 149 Photochemical reactions, 105
Lanthanide luminescence, 40 Photocycloaddition, 126
Lanthanide MOFs, 64, 136 Photodimerization, 110
Ligands, chiral, 149 Photoluminescence, 21
Ligand-tometal charge transfer (LMCT), Photopolymerization, 139
41, 46 Photoreactivity, 105, 112, 139
Lithium niobate (LiNbO3), 147 Photoreduction, 91
Luminescence, 2, 27, 125, 167 Photosensitizers, 91
Index 189

Pillared-layer MOFs, 133 Tetrakis[4-(carboxyphenyl)oxamethyl]


Porous coordination polymers (PCPs), methane acid, 96
1, 167, 169 Tetrakis(imidazole), 33
Potassium dihydrogen phosphate (KDP), 147 Tetrakis(imidazol-1-ylmethyl)methane, 96
Potassium titanyl phosphate (KTP), 147 Tetramethyl-4,4-bipyrazolate, 15
Push–pull effect, 145, 150 Tetramethyl-4,4-difluoro-8-bromomethyl-4-
Pyridine-2,6-dicarboxylic acid, 178 bora-3,4-diaza-sindacene, 76
Pyridyldicarboxylate ligands, 7 Thiophenyl-derivatized carboxylic acid, 42
Pyrizine-2,3-dicarboxylic acid, 11 Thrombin, 182
Transformations, photocatalytic, 97
Triazolates, 3
Q Triazole, 159
Quantum dots (QDs), 47, 173 Trifluoroacetates, 110
Trinitrophenol, 181
Tris(4-carboxyphenyl)benzene, 61
R
Ratiometric sensors, 71
Reticular Chemistry Structure Resource V
(RCSR), 3 Vapor sensors, 69

S W
Second harmonic generation (SHG), 145 Water splitting, 89
Secondary building units (SBUs), 33 White-light-emitting materials, 60
Small molecules, sensing, 179 White-light luminescence, 27
Smart materials, 2
Solid-phase microextraction (SPME), 19
Solid-state polymerization, 139 X
Solid-state reactions, 105 X3B, 96
Structural transformation, 105, 167, 176
trans-4-Styryl-1-methylpyridiniumiodide, 175
4-Styrylpyridine, 119 Y
Syntheses, 32 Ytterbium MOFs, 66

T Z
Terephthalic acid, 96 Zeolites, 2, 93
Tetraazacyclotetradecanetetrapropionate, 178 Zeolitic imidazole frameworks (ZIFs), 34
Tetrakis( p-benzoic acid)pyrene, 68, 179 Zr-carboxylate MOFs, 92
Tetrakis(4-carboxyphenyl)ethylene (TCPE), 40 Zwitterionic metal complexes, 110

You might also like